Вы находитесь на странице: 1из 57

State of the art report: Analysis and design

Analyse et conception : un état de l’art

B. Simpson & P. Morrison


Arup Geotechnics, United Kingdom
S. Yasuda
Tokyo Denki University, Japan
B. Townsend
Hatch Mott MacDonald, LLC, USA
G. Gazetas
National Technical University of Athens, Greece

ABSTRACT
This paper is concerned with analysis and design across the breadth of geotechnical engineering. Sections by individual experts relate
to Codes and standards, Deep foundations, Embankments and slopes, Underground construction and Seismic design. Each considers
recent developments, including economy and sustainability. Recent geotechnical codes and standards from Europe, North America
and Japan are compared. All adopt forms of limit state design and address familiar geotechnical problems of uncertain ground
behaviour and complex interaction of loads with frictional materials. Some topics which are still under debate are discussed. Design
of deep foundations is a blend of empiricism and theory. Developments including drilled and grouted piles in rock and offshore
driven piles are presented. Topics considered include assessment of ultimate capacity and the environmental benefits of pile re-use
and piles as heat transfer elements. Design of embankments and slopes depends on empirical rules and observation, as well as theory.
In Japan, tolerance to displacements following earthquakes is a particular consideration. Embankments used as transport corridors or
as building platforms are discussed. Failures involving uncontrolled fills are particularly noted. For construction underground,
understanding of the fundamental physics is imperative, along with comprehensive process of checking during both design and
construction. Greater use of cement replacements would aid sustainability. The final section of the paper considers performance-
based design of foundations against two seismic hazards: emergence of a rupture underneath a structure, and bearing capacity
mechanisms for slender structures on shallow foundations.
RÉSUMÉ
L’article étudie l’analyse et la conception d’ouvrages géotechniques variés. Les différentes sections abordent les codes et standards,
les fondations profondes, les pentes et talus, les constructions souterraines, puis le génie parasismique. Les développements récents,
notamment en matière d’économie et de développement durable, sont présentés. Les codes et standards utilisés actuellement en
Europe, aux Etats-Unis, et au Japon sont comparés. Tous adoptent des formes de conception aux états limites et abordent les
problèmes courants causés par les incertitudes sur le comportement du sol, et par les interactions structures-milieux frictionnels.
Certains sujets encore en débat sont abordés. La conception de fondations profondes mélange empirisme et théorie. Plusieurs
développements sont présentés, tels les pieux forés-cimentés dans les roches et les pieux battus en mer. Estimation de la capacité
portante ultime, intérêt environnemental de la réutilisation de pieux, et utilisation des pieux pour transférer la chaleur sont, entre
autres, considérés. La conception des pentes et talus repose sur des lois empiriques, l’observation, et la théorie. Au Japon, la tolérance
aux déplacements dus aux séismes est d’une importance particulière. L’utilisation des talus comme couloirs de transport, ou comme
plateformes pour les constructions, est abordée. Les ruptures de remblais non contrôlés sont aussi étudiées. Les constructions
souterraines exigent une bonne compréhension de la physique fondamentale et la vérification des phases de conception et de
construction. Une utilisation plus fréquente de substituts au ciment serait plus environnementale. La dernière section concerne la
conception des fondations contre deux phénomènes liés aux séismes: apparition de discontinuités sous la structure, et modes de
ruptures par perte de capacité portante des fondations superficielles de constructions élancées.
Keywords design, analysis, codes, standards, piling, tunnelling, slopes, embankments, seismic

1 INTRODUCTION the design process. Analysis is one part of design, involving


calculations carried out by engineers, with or without the aid of
The president of the ISSMGE, Professor Pedro Seco e Pinto, computers. The main thrust of the contributions relates to
invited an international group of experts to prepare this State of design aspects. References are made to supporting analysis, and
the Art Report on Analysis and Design in geotechnical it is assumed that advanced techniques such as finite element
engineering. It was agreed that five topics would be included in computations will be available and used when appropriate.
the report: Codes and standards, Deep foundations, Slopes and
embankments, Underground construction and Seismic design. 2 CODES AND STANDARDS
Each author was asked to provide a brief review of the
fundamentals of technical understanding of most importance to
their topic and to comment in particular on recent 2.1 Introduction
developments, especially since previous state-of-the-art papers
on the topic. Comments are provided on what goes wrong, and Codes and standards aim to set out the process and procedures
emphasis is given to questions of economy and sustainability. of design, or at least to identify the basic ingredients and limits
Design is decision. In the context of this paper, design is the of acceptable good practice. Figure 2.1 shows the functions
complete process or sequence of decisions which determine performed by codes in many projects, providing the
what is actually built. The decisions may be taken by clients, communication between society, clients, data and analysis in the
consultants, contractors, site workers or others – all are part of process of design.

Proceedings of the 17th International Conference on Soil Mechanics and Geotechnical Engineering 2873
M. Hamza et al. (Eds.)
© 2009 IOS Press.
doi:10.3233/978-1-60750-031-5-2873
2874 B. Simpson et al. / State of the art Report: Analysis and Design

Designs Grounded on a Performance-based Design Concept


(Japanese Geotechnical Society 2006).
The Eurocodes use the term “actions” in place of the term
“loads” used in American and older British codes. English
translations of Japanese documents use both terms. In this
paper, the term “loads” will generally be adopted, except in
direct quotations.
Clause numbers from codes are shown in brackets thus: { }.

2.2 Limit states and working states

Investigators of failures frequently report that the main cause


was some onerous condition which had not been considered in
the design. In contrast, it is often noted that failures were not
caused by normal or even somewhat excessive statistical
Figure 2.1: Communication facilitated by codes variations of parameters which had been properly reviewed and
for which standard margins of safety had been allowed. It
In this paper, the term “code” will be used in a broad sense. appears to be important, therefore, that designers’ attention is
It includes “design standards” such as the Eurocodes, “design drawn towards a reasonable range of extreme situations against
codes” such as Japanese codes, and “design specifications” such which to check each design. This is the underlying philosophy
as the AASHTO LRDF Bridge Design Specifications. The of limit state design.
older style of British “code of practice”, which often included a Limit states are “states beyond which the structure no longer
more discursive account of the geotechnical design process satisfies the relevant design criteria”. That is, the structure, or
rather than definitive rules, is within the scope of this part of it, is on the point of “going wrong” – failing in a sense
discussion, but specifications for construction are outside the that would be understood by a client, user or non-technical
scope. observer. All the codes place the limit state on the safe side of
the failure, ie when the structure is almost failing but has not yet
Part of the function of a code is to provide a link between failed. Limit states are classified into various types, used
analysis and design, showing how analysis can be used as a tool selectively in the codes.
in the design process. Codes may provide limits on the analysis In the Eurocodes, Ultimate limit states (ULS) are “associated
methods to be used, deliberate or unintended, and generally with collapse, or with other similar forms of structural failure”.
provide safety factors or margins, establishing gaps between They generally relate to danger or severe economic loss.
what is analysed, failure states and what is most likely to occur Serviceability limit states (SLS) “correspond to conditions
in practice. beyond which specified service requirements for a structure or
Society requires safety and serviceability. Codes provide a structural member are no longer met”. These are unwanted but
link between the activities of the designer and the requirements more manageable if they occur, allowing the structure to remain
of society. They are the point at which society’s qualitative in service if necessary repairs are carried out. They may be
requirements that structures should be at least as reliable as they subdivided into irreversible SLS (eg cracking) where the
traditionally have been are interpreted in the forms of consequences of exceedance remain after the disturbing action
procedures and numbers. Society, or at least clients, also is removed and reversible SLS (eg unpleasant vibration) where
requires economy, to be balanced with safety and serviceability. the consequences do not remain. The Eurocodes also consider
Although codes have often promoted safety and serviceability accidental design situations, though these are regarded as
with less regard for economy, they also have a responsibility potential causes of limit states, rather than limit states in
here, especially in a world in which limitations of resources are themselves.
increasingly recognised. The Japanese Geocode 21 emphasises that the ULS is still on
Perhaps the greatest challenge for drafters of geotechnical the safe side of collapse: “the structure may have sustained
codes is to find the best balance between fixed rules and considerable damage, but not to the extent that the structure has
personal expertise, both of which are essential to successful reached failure, become unstable, collapsed, or to the extent that
design. Geotechnical engineering relies heavily on the would result in serious injury or the loss of life.” Similarly, at
knowledge and judgement of individual designers. Inevitably SLS there is to be no effect on durability and “regular use of the
this is somewhat subjective, depending on the training and structure is possible, without repairs.” This code also has a
experience of each individual, but equally it is indispensably Reparability limit state “in which damage to the structure has
valuable, especially when shared in discussion with others; occurred and may have affected the durability of the structure.
codes must allow and encourage this. For both individuals and However, regular use of the structure is possible to a limited
the profession as a whole, past mistakes are a major source of extent and there are reasonable prospects for full functionality
learning, and both must aim to avoid repetition of these. Codes of the structure if economically-feasible repairs are performed.
are an important part of the profession’s corporate memory, This limit state can be interpreted as the state in which the
helping to assure society that past mistakes will not be repeated. majority of the value of the structure has been preserved. In
Equally, though, modern codes are aiming to pull professional addition, the reparability limit state sometimes implies a state in
practice forward, closer to the best available “state of the art”. which marginal use of the structure is possible for rescue
In the following sections, approaches currently taken by operations immediately following an extraordinary event such
codes to geotechnical analysis and design will be reviewed and as a large earthquake.”
compared. Particular emphasis will be given to European, The AASHTO Bridge Code has Strength limit states
North American and Japanese developments, noting both the “relating to strength and stability during the design life”. This
broad areas of agreement and particular points of debate within is a more technical definition than the Eurocodes’ ULS, but
and between these communities. Specifically, the main from most points of view it serves a similar purpose. The same
documents included in the review are the Eurocodes, code has Service limit states “relating to stress, deformation,
particularly Eurocode 7 – EN1997-1 (2004), Geotechnical and cracking under regular operating conditions”. AASHTO
design; the AASHTO LRFD Bridge Design Specifications also have Extreme event limit states “relating to events such as
(2008); and Japanese Geocode 21: Principles for Foundation earthquakes, ice load, and vehicle and vessel collision, with
return periods in excess of the design life of the bridge.” These
B. Simpson et al. / State of the art Report: Analysis and Design 2875

may be compared with the Eurocodes’ accidental design safety margins by a combination of two features: somewhat
situations noted above. cautious characteristic values, determined by the designer in the
Offshore codes such as Det Norske Veritas (1993) include case of soil properties, and the application of partial factors,
other “limit states” such as progressive collapse limit state, prescribed by the codes on behalf of society. Usually,
fatigue limit state, though these may be considered more “cautious” values of strength are lower than the most probable,
descriptive of the way the limit state is reached than of the limit but for some limit states values on the high side are more
state itself. The survivability limit state, however, generally critical.
does describe a particular state of the structure. In the case of loading, the values themselves are largely
In Limit State Design, attention is directed to states at or prescribed by the loading codes. For manufactured materials,
close to failure, which it is hoped will not occur in practice. characteristic values of strength are derived as a fractile of the
The alternative is Working State Design in which attention is test results of a closely specified type of test, typically 2
directed to the expected, desired state in which the structure is standard deviations from the mean, giving a frequency of about
performing successfully under its expected loading. Its safety is 2.3% in the tests. It is commonly the case that code drafters are
then usually checked by requiring that the degree to which more knowledgeable about both loading and structural material
material strengths or ground resistances are mobilised is limited. properties than are designers. For ground material the situation
Limit state design codes usually require that at least two is more complex, however, for several reasons: (a) ground
limit states be checked. One of them is normally a materials cover a wide range of types, some more easily
serviceability limit state, for which unfactored parameter values sampled than others, so a range of testing approaches is
are usually used in calculations. This is the nearest approach to unavoidably adopted; (b) in most, but not all cases, the
an analysis of the expected performance of the structure, but it parameter relevant to the design is close to a mean value,
is more pessimistic than is really expected, as discussed further averaged over a large zone or surface within the ground, rather
below. The other limit states involve more severe situations, than to an extreme derived from testing small samples; (c) the
often derived by applying partial factors to selected parameters uncertainty and variability of the ground on a particular site is
representing loads, ground or structural material strengths, often much better known to the designer than it could be to the
ground or structural resistances, and possibly other parameters code drafter; (d) the process of construction sometimes changes
including water pressure. the ground properties; (e) data and observations obtained from
Despite some differences of emphasis, or of drafting, the similar projects may be highly relevant to selection of
developers of the modern codes agree that a limit state appropriate parameter values; (f) full scale testing is quite often
approach, concentrating on failure, or near failure, is to be incorporated into the design process, particularly for piles and
preferred (see, for example, Becker, 1996). The present authors ground anchors.
agree with this conclusion, adding the following observations: This problem is not new: in past practice it was usually
1. Limit state design implies that the states studied should not unclear how conservative the code drafters assumed the values
occur. For ULS, their probability is to be extremely low, so of parameters to be. An early attempt at providing more
calculations are directed to a virtually non-existent state. guidance was represented by CIRIA Report 104 (Padfield and
This may confuse designers who more readily imagine Mair 1984) which asked users to assess either “moderately
states which they consider fairly likely to exist. In this conservative” or “worst credible” values for soil strength
respect, working state design is easier to understand. parameters and provided differing factors for these two cases.
2. It was noted above that actual failures are often caused by The same approach is taken in the more recent replacement of
the occurrence of situations that had not been properly this report, CIRIA Report C580 (Gaba et al 2003).
foreseen during the design process. Some of these might be The developers of various geotechnical codes have taken
accepted as virtually unforeseeable, while others are errors different approaches to this problem. After very lengthy
of process on the part of the designer, such as misjudgement debates in Eurocode 7, the basic definition of characteristic soil
of soil properties or failure to check a critical mechanism. It parameters was agreed as: “a cautious estimate of the value
is essential, therefore, that codes continue to emphasise affecting the occurrence of the limit state”. Additional text
good geotechnical practice and process, avoiding too great a makes it clear that this estimate is to include allowance for the
concentration or reliance on factors of safety. extent of ground involved in the limit state, and hence average
3. Calculation errors are discovered very often during effects over that extent, and the effect of construction activities
investigations of failures. Complexity in codes seems likely on ground properties. In making this estimate, input from
to increase the number of errors, so it is important, as a laboratory and field tests, and also well-established experience
safety issue, that code provisions are kept as simple as are required, but the type of testing is not restricted and a
possible. distinction between situations in which spatially averaged
4. A few of the calculation errors prove to be critical, but it values or extreme values govern is encouraged. It is suggested
seems likely that many others are covered by the margins of that if statistical approaches are applied to the data “the
safety traditionally in use. Human error is therefore a characteristic value should be derived such that the calculated
significant uncertainty that must be included in any rational probability of a worse value governing the occurrence of the
attempt to determine factors of safety. Failing this, overall limit state under consideration is not greater than 5%”. Thus the
factors of safety can only be reduced with the utmost emphasis throughout is on engineering assessment of the values
caution. actually governing behaviour in the ground, which might differ
5. The aim, however, should be to reduce factors of safety, and from those measured in tests, with a requirement that the
certainly not to increase them, in general. Unnecessarily assessment be “cautious”, not a best estimate of the most likely
large factors of safety lead to increase in cost, and wastage value, especially if there is considerable uncertainty. It is
of materials and energy. implied that “cautious” means that the chance that values worse
than the selected value will actually occur in such a way as to be
2.3 Assessment of parameter values – characteristic values the cause of a real problem (a limit state) is about 5%. The
expertise of the engineer is by no means relegated, but is
Prescribed values for partial factors provided by codes have incorporated into the assessment of the characteristic values.
little meaning unless it is clear how the parameter values are The present authors would argue that this is no different from
initially to be selected before being factored. These unfactored previous good practice, and essentially the same as the
parameters are referred to as characteristic values of loads, “moderately conservative” values of CIRIA Report 104;
strengths or resistances in Eurocodes and Geocode 21, and as engineers have generally aimed for a mean value, but have been
nominal values in AASHTO. In effect, modern codes provide
2876 B. Simpson et al. / State of the art Report: Analysis and Design

slightly cautious, especially when the degree of uncertainty was value; the Japanese position might be slightly different. The
significant. recommendations of Becker and of Foye et al, intended for
It is important to recognise that in a typical situation North American practice, encourage the adoption of a limited
involving large amounts of ground a “probability of a worse range of specific tests, CPT and SPT, with the characteristic
value governing the occurrence of the limit state under value derived directly from the test results (including
consideration is not greater than 5%” is very different from a modifications for stress level and possible removal of rogue
5% fractile of test results. Schneider (1997) discussed situations values). This would largely remove judgement from the
where the zone of ground affected is large enough that its process, although considerable geological judgement remains in
behaviour will be close to the overall average of the soil zoning the site, both in plan and in level. In contrast, the
considered. He argued that a “cautious estimate” with a 5% European approach incorporates the designer’s expertise into
chance of a worse overall value would lie about 0.5 standard the assessment of characteristic values. It allows assimilation of
deviations from the mean of relevant measured values, adjusted results from a range of test types and sources, and encourages
as necessary for sampling disturbance and construction effects. engineers to consider the relevance of the test results to the way
The marked difference between these two assessments can be the ground will respond at a specific limit state, including
seen in Figure 2.2. effects of construction on its behaviour. Eurocode 7 also notes
the need to be aware that extreme values of strength may be the
critical ones if a failure mode could exploit them.
An alternative approach is also mentioned by Becker (1996)
and has been used in Swedish practice (Boverket, 1995). In this
approach, the “Characteristic value” is taken to be a simple
mean value, or best estimate of the actual value. The value of
applied partial factors is varied in a prescribed manner related to
the assessed reliability and variability of the data from which
the mean was derived, and its likely ductility. The author
understands that Swedish codes are changing towards a
definition which includes some of the uncertainty in the choice
of characteristic value, and is therefore closer to the Eurocode
definition.
With relatively small differences in approach, all these
documents share the aim of obtaining the most useful, rational
and transparent combination of objective test results with
Figure 2.2: Alternative derivations of “characteristic values” engineering knowledge and experience, necessarily subjective
to some extent. A purely objective process would disregard
A similar proposal was made by Dahlberg and Ronold large amounts of human knowledge which are essential to
(1993) for design of offshore foundations and recommended by geotechnical design.
Becker (1996) for more general use. This involves the use of a
“conservatively assessed mean” (CAM) as the characteristic 2.4 Can the various limit states really be treated separately?
value, also about 0.5 standard deviations from the mean of the
test results. These authors state that for a normal distribution In principle, limit state design requires separate consideration
75% of the measured values would be expected to exceed this and analysis of various different limit states, for example,
value. (More accurately, this requires an offset of 0.69 standard ultimate (or strength) and serviceability limit states. Separation
deviations from the mean, for a normal distribution, as shown in makes it easier to specify the requirements of design for each
Figure 2.2). Foye et al (2006b) take up the same idea proposing limit state, and many authors, eg Becker (1996), have argued
to use a CAM with 80% exceedance, equivalent to 0.84 that this is essential for the “rational” derivation of values for
standard deviations below the mean of the test results for a partial factors, ie on a probabilistic basis. The split between
normal distribution. These proposals are made in the ultimate and serviceability has encouraged development of
development of North American practice, though at present the testing methods and analytical tools for serviceability
AASHTO Specifications do not use the term “characteristic” calculations, generally aiming to provide ground stiffnesses for
but refer less specifically to nominal values related to calculation of displacements. The more detailed prescriptions
“permissible stresses, deformations, or specified strength of of the codes have generally been dominantly ULS, partly
materials”. because SLS criteria are the province of the client rather than
Japanese Geocode 21 {2.4.3} defines characteristic values of being requirements of society.
geotechnical parameters as “the representative values that have Parameter values and related calculation models giving
been cautiously estimated as the most appropriate values for the reliable predictions of displacement are often very difficult to
foundation-ground models in order to predict the limit states obtain. In contrast, those for mechanisms of plastic failure are
checked in the design”; this is apparently close to the Eurocode generally easier and more reliable; for example, the angle of
definition. It continues: “The characteristic value of a friction of a soil is much easier to obtain than its deformation
geotechnical parameter is principally thought of as the average properties. Because of this, Eurocode 7 also allows SLS to be
(expected value) of the derived values”, but it adds covered by the strength-based (plastic mechanism) calculations
qualifications to this statement that diverge away from a simple in some cases, as shown by the following extracts.
statistical average, and shows the same in a mathematical {2.4.1(4)P} If no reliable calculation model is available for a
formula. The terms “cautiously estimated” and “expected” are specific limit state, analysis of another limit state shall be
potentially in conflict, but the overall impression given is that carried out using factors to ensure that exceeding the specific
the characteristic value may be somewhat more conservative limit state considered is sufficiently improbable. Alternatively,
than the statistical mean. However, commenting on Geocode design by prescriptive measures, experimental models and load
21, Honjo et al (2005) particularly emphasise that “the tests, or the observational method, shall be performed.
characteristic value is defined as a mean value of a geotechnical {2.4.8(4)} It may be verified that a sufficiently low fraction
parameter”. of the ground strength is mobilised to keep deformations within
Thus there seems to be approximate agreement between the the required serviceability limits, provided this simplified
Eurocodes and the North American proposals about the level of approach is restricted to design situations where:
conservatism most usefully associated with the characteristic
B. Simpson et al. / State of the art Report: Analysis and Design 2877

• a value of the deformation is not required to check the structural design in situations where unfavourable and
serviceability limit state; favourable loads tend to cancel in the expected state, leaving
• established comparable experience exists with similar only small working stresses, but where a slight increase in the
ground, structures and application method. unfavourable loads could lead to proportionately very much
{6.6.2(16) [for spread foundations]} For conventional bigger increases of stress.
structures founded on clays, the ratio of the bearing capacity of In November 1965, during severe wind conditions, three of
the ground, at its initial undrained shear strength, to the applied the eight cooling tower shells collapsed at the Ferrybridge C
serviceability loading should be calculated (see 2.4.8(4)). If this Power Station in Yorkshire, UK (Figure 2.3). The Committee
ratio is less than 3, calculations of settlements should always be of Inquiry into the collapse reported that wind conditions were
undertaken. If the ratio is less than 2, the calculations should “very considerably lower than the probable maximum values”
take account of non-linear stiffness effects in the ground. and that the structural analysis had been carried out correctly.
{7.6.4.1(1)NOTE [for pile design]} For piles bearing in They reported: “The important membrane stresses in the shells
medium-to-dense soils and for tension piles, the safety of the towers are the resultant of compressive stresses due to
requirements for the ultimate limit state design are normally dead weight and tensile stresses due to wind uplift: the resultant
sufficient to prevent a serviceability limit state in the supported tensile stress, being the difference between two large quantities,
structure. is sensitive to small variations in the wind loadings.” This
The first of these extracts refers to limit states in general. event was critical in the rejection of working state design and
The second and third occur in sections on SLS, allowing development of load factoring, indicating the importance of
strength calculations which may be different from those considering each independent load separately.
required for ULS to be used to demonstrate limited settlement. Ferrybridge was not a geotechnical failure, but all the loads
It will normally be the case that this SLS condition is, by in any structure have eventually to be brought down to the
inspection, more severe than the ULS condition which would ground, so the same thinking applies to foundation and
use the same type of calculation, so one calculation checks both geotechnical designs. Schuppener et al (2009) discuss the
SLS and ULS. The fourth extract, referring to piles, specifically examples shown in Figure 2.4. In each case, the forces
states that the ULS calculations will normally cover SLS also; transmitted to the ground may differ appreciably from the most
again, one calculation covers both. In the European system, likely working state due to relatively small changes in the
actual values for partial factors are set by each nation, so each structure and its loading. In Figure 2.4a, small changes in
must ensure that the values set comply with this note; the loading change the forces in the piles from tension to
application of this in developing the UK National Annex for compression. In Figure 2.4b, small changes determine whether
EC7 (BSI 2007) is discussed by Bond and Simpson (2009). an anchor is needed or not at point A.
If the values of partial factors were derived by probabilistic To avoid problems of this type, the drafters of modern codes
analysis, it would be confusing to mix SLS and ULS, as noted have taken the view, rightly in the opinion of the present
by Becker (1996). However, in practice the values adopted in authors, that designers should be required to consider states in
codes to date are mainly chosen to reproduce existing which relatively extreme values of parameters take the structure
experience of successful behaviour. Since this implies success close to failure.
both in ULS and SLS, it is clear that the factors being chosen
are adequate in most cases to provide for both, and it may be
very difficult to determine which of the two limit states actually
governs the factor values. Factors based on past successes
apparently cover all significant variables, to a degree which is
generally adequate. Besides uncertainly of loads and material
strengths, this includes calculation models, geometric
uncertainties and human errors.
In the opinion of the authors, statistical studies are useful to
code drafters in helping to allocate partial factors in appropriate
proportions between the parameters. However, when the
overall level of safety derived by modern codes is fixed by
calibration with past successful experience, actual magnitudes
may well cover more than one limit state and inevitably allow
for a greater range of uncertainties than the statistical variation
of the basic parameters. A lack of theoretical purity in approach
is probably of little concern to designers, provided codes are (a)
clear about what limit states are covered by each of their
requirements, and in particular when it is necessary to carry out
specific calculations of deformation and displacement.

2.5 Review of the alternative approaches

2.5.1 Working state design


As noted above, in Working state design attention is directed to
the expected, desired state in which the structure is performing
successfully under its expected loading. Its safety is then
usually checked by requiring that the degree to which material
strengths or ground resistances are mobilised is limited. This
means that “permissible stresses” are not exceeded, so the
approach is also called Permissible stress design or Working
stress design (WSD).
Working state design has the advantage that the designer is
asked to consider states which are easier to imagine because (b)
they are likely to occur. Its weaknesses became apparent in Figure 2.3: Collapse of Ferrybridge cooling towers
2878 B. Simpson et al. / State of the art Report: Analysis and Design

robust and reasonably simple to apply. In the author’s


experience, the greatest complexities and lengthy debates arise
in design practice when the system proposed leads to obviously
unreasonable results; sometimes this occurs as a result of over-
enthusiasm for simplicity.

2.5.3 Factors applied to material strength or resistances


On the resistance side, the main debate is whether to apply
factors γM to material strength (cu, c′, tanφ′) or factors γR or φ to
derived resistances such as bearing capacity or lateral
resistance. Becker (1996) put the case for resistance factors as
follows:
“Although the factored strength approach may be considered
as being more elegant and sophisticated, the factored overall
resistance approach has a significant advantage over the
factored strength approach in that the derived resistance factor
reflects not only uncertainty in strength but also uncertainties
associated with the analytical models, site conditions,
(a)
construction tolerances, and failure mechanisms. The factored
strength approach alone does not capture all sources of
uncertainty in the calculation of resistance. There is a need to
also take into account the uncertainties stated above and others
such as the development of excess pore-water pressure and
stress-strain behaviour. The factored strength approach also
(b) does not capture the true mechanism of failure when failure is
influenced by nonlinear soil behaviour. … The factored
Figure 2.4: Examples of balanced loads (after Schuppener et al 2009) resistance approach is similar to WSD and may be viewed as a
logical extension to WSD. Therefore, it would be familiar to
2.5.2 Introduction of partial factors and, hence, better received by geotechnical engineers, which
For ultimate or strength limit states, the requirements can would allow for a smoother transition from WSD to LSD for
generally be resolved into a condition of the form Ed ≤ Rd. That foundation design.”
is, the design effect of loads must not exceed the design In Eurocode 7, both possibilities are allowed and the choice
resistance, the term “design” implying that all necessary factors is left to individual nations. There has been no general
are already incorporated. This condition may be checked at agreement as to which approach is preferable or, indeed,
many different points in a structure, typically for example as a whether either approach is best in all circumstances. The main
thrust, tension, shear force or bending moment in a structural argument voiced in favour of the resistance factor approach has
member, or as a bearing pressure, disturbing moment or lateral been the second main point made above by Becker, that this
resistance in a geotechnical calculation. The term Ed is formed more readily reflects existing practice based on single factors of
by the combination of separate loads with their own factors and safety. However, some European countries diverged from that
the term Rd is calculated from characteristic values of ground system many years ago; examples include CIRIA Report 104
strength using theoretical or empirical relationships of varying (Padfied and Mair 1984), the Danish code (Dansk
provenance and reliability. Significant differences of opinion Ingeniorforening 1984), Norwegian codes (Det Norske Veritas
exist in deciding where in the process of deriving Ed and Rd the 1993), and the British retaining walls code (BS8002 1994). It
factors should be applied. would be difficult to imagine a single set of factors which could
Eurocode 7 effectively allows a very general equation: adequately take account of “uncertainties associated with the
analytical models, site conditions, construction tolerances,
γE E{γF Frep; Xk/γM; ad} = Ed (2.1) failure mechanisms, … excess pore-water pressure and stress-
≤ Rd = R{γF Frep; Xk/γM; ad}/γR strain behaviour”. The Eurocode approach is generally not
specific about which analytical models are to be used but
requires that they should be “accurate or erring on the side of
in which the main variables are Frep, Xk and ad – representative safety”; where necessary, an additional resistance factor (or
loads (or “actions”), characteristic ground strengths and design “model factor”) may be introduced to ensure this. The code
geometrical parameters (representative actions are derived from requires other uncertainties such as construction tolerances to be
characteristic actions using load combination factors, outside considered specifically if they are significant and similarly
the scope of this paper). The functions E{} and R{} represent requires either that extreme pore pressures are considered or
the derivation of the load effect from the individual loads and of that load factors are applied directly to them. In all systems,
the corresponding resistance from the ground strengths and minor variations in these secondary variables are implicitly
other parameters. The factors are γE, γF, γM and γR on load covered by the factors on the leading variables.
effects, individual loads, material strengths and resistances, Simpson (2007) considered some of the advantages of a
respectively, some of which may be set to unity. The material material strength approach in geotechnical design, in particular:
strength term Xk is included in the load effect side of the • It facilitates consistent analysis of combined problems,
condition because ground strength may affect load effects such which are very common, involving, for example, a slope,
as lateral earth pressures; similarly the load term Frep is included loaded by a structure, supported by a retaining wall, itself
on the strength side because in frictional soils loads may affect supported by anchors and foundations. An example is
strength – for example normal loads enhancing sliding shown in Figure 2.5.
resistance. In North American usage, the reduction factor φR is • Because the strength of soil is derived from friction, non-
used in place of 1/γR. linear, or disproportionate relationships between soil
Between the various groups working on code development, parameters and resistances are common. In these
opinions have diverged about which of the factors γ can most circumstances, it is important to check designs with factors
usefully be set to unity, their effect being incorporated into of safety applied to the basic strength parameters of the soil,
other factors. All developers aim to provide a system which is as explained fully in EN1990 {6.3.2}. Such non-linear
B. Simpson et al. / State of the art Report: Analysis and Design 2879

also shows the equivalent value of φR provided by a constant


factor of 1.25 applied to tanϕ′, a typical value of γϕ used in
EC7; adjustments have been made to allow for the differences
between load factors in the two approaches. (The plot for DA2
will be discussed later.) It can be seen that the constant γϕ
provides an equivalent φR having a similar variation with ϕ′ to
that proposed by Foye et al, and corresponds, according to their
analysis, to a reliability index β of between 2 and 2.5. The
precise value of β is dependent on the way the characteristic or
nominal ϕ′ is derived; Foye et al show different values when it
is derived from SPTs because greater uncertainty is involved.

Figure 2.5: Combined geotechnical and structural design situation

relationships occur, for example, in the derivation of


bearing capacity or of passive resistance from angle of
shearing resistance, as illustrated in Figure 2.6. (Kp is
2
plotted for δ/φ= ⁄3.)
• It can be used readily with both simple hand calculations
and more complex finite element calculations. Introduction
of resistance factors into finite element computations, which
essentially require overall equilibrium, has proved to be
difficult. In the authors’ view, demonstration of equilibrium
of complete systems is fundamental to good design practice.

Figure 2.7: Relation of φR to ϕ recommended by Foye et al (2006b)

The implications of this comparison for ULS can be seen in


Figure 2.8, which shows results obtained by Foye et al (2006b)
of the width of strip footings as a function of nominal angle of
shearing resistance ϕ′, obtained from interpretation of CPT
results. Also shown are calculations following EC7 DA1,
plotted against characteristic angle of shearing resistance. IN
both cases, the footing with B has been normalised in relation
to the soil weight density γ and the total design load ΣFd used by
Foye et al. They use a formula for the bearing capacity factor
Nγ taken from Salgado et al (2004) which is more conservative
than the EC7 formula, so calculations using the EC7 partial
factors are shown for both bearing capacity formulae in Figure
2.8. As might be anticipated from Figure 2.7, the two sets of
results are similar EC7 results are similar, with the EC7 results
somewhat less conservative.
Figure 2.6: Non-linear relationships between material property and An interesting feature of this work is that Foye et al based
resistance. their analysis on an assessed coefficient of variation of angle of
shearing resistance, which would most readily be represented by
Foye et al (2006 a,b) report studies intended to derive values a factor on tanϕ′ (or ϕ′ - it makes little difference), rather than
for resistance factors φR for design of spread foundations. Their by a φR. The same implication was made in previous work by
work is based on statistical studies of the variability of material Burland et al (1981) who developed a new method for
strengths, as measured by CPT and SPT, and does not representing factor of safety in design of embedded retaining
incorporate explicitly the other variables listed by Becker walls. Again, they justified it by showing that it was equivalent
(1996). Indeed, these would be very difficult to include since to a constant factor on soil strength, implying indirectly that
statistical studies of the behaviour of spread foundations in factoring soil strength is a sound way to proceed. It seems that
practice are not available. Starting from an assessment of the geotechnical engineers intuitively expect that a constant factor
coefficient of variation of angle of shearing resistance, ϕ′, they on tanϕ′ is a good way to give a reasonably constant level of
aim to achieve constant reliability for bearing capacity and safety. This may reflect the point noted above that when
conclude that ideally the factor φR on bearing capacity should strength is derived from friction, non-linear, or disproportionate
vary as a function of ϕ′. Figure 2.7 shows how their relationships between soil parameters and resistances are of
recommended value of φR varies with “nominal” (equivalent to concern.
“characteristic”) value of ϕ′ derived from CPT results. The One problem of a material strength approach is that some
graph is plotted for a live/dead load ratio of 0.5, and four curves calculation methods derive bearing resistance, for example,
are plotted for four values of reliability index β that they directly from results of in situ tests, without the use of a
consider to be relevant, with preference for β=3.0. The figure
2880 B. Simpson et al. / State of the art Report: Analysis and Design

combinations” The Specifications state “Bearing resistance


shall be investigated at the strength limit state using factored
loads” {11.6.3.2}. Despite this, for the calculation of load
inclination unfactored loads are used {eg 10.6.3.1.2a}; an
earlier edition gave the explanation “To preserve the angle of
inclination of the resultant, unfactored loads are used. When
factored loads are used, the failure surface beneath the footing
could be different than the one due to the applied loads, and the
result may become overly conservative.” Modjeski and Masters
Inc (2003) provide an example of this for a spread foundation
for a retaining wall designed to AASHTO Specifications.
The various National Annexes to EC7 take differing views
of this question, as described below.

2.6 Eurocode 7

EC7 requires that both ULS and SLS be considered. Most of its
text refers to ULS; text referring to SLS was discussed above.
For ULS, the main approach is based on use of partial factors,
but opinions in Europe differ about where and how these should
be applied. This is left to national choice, and the values to be
adopted for partial factors may also be varied nationally. Three
Figure 2.8: Footing dimensions according to Foye et al (2006b) and alternative “Design Approaches” have been developed,
EC7 combining partial factors in different ways; the factor values
proposed in the European document are shown in Figure 2.9.
parameter of material strength. If it is significant, this difficulty Bond and Harris (2008) have summarised the adoption of the
could be overcome by having a resistance factor which varies as three Design Approaches by the various countries in Europe, as
a function of the test result, much as suggested by Foye et al. shown in Figure 2.10, for designs other than slope stability.
The special case of pile design is considered below. Most countries that have adopted DA2 for other purposes have
decided to use DA3 for slope stability.
It was noted above that because factor values are inevitably In Design Approach 1 (DA1), two “combinations” of partial
based on experience to a considerable extent it is impossible to factors are specified, and the design must be shown to
distinguish whether they are required by ULS or SLS accommodate both combinations (Figure 2.9). Essentially, they
considerations. Bolton (1993) has argued that in many cases are used in the same way as load combinations, but the concept
SLS requirements can reasonably be accommodated by is extended to include material strengths and resistances. Partial
checking the degree of mobilisation of material strength (also factors are generally applied to either loads (before
Osman et al 2004, Osman and Bolton 2006). This is also combination) or ground strengths (before calculation of
readily facilitated by factors on soil strength, relieving code resistances), though with some exceptions. In countries that use
drafters, to some extent, of the need to make the difficult DA1, the factors on ground materials and strengths are
distinction between requirements for ULS and SLS. generally set to 1.0 in Combination 1; the factors adopted in the
United Kingdom are shown in Figure 2.11. For design of piles
2.5.4 Factors applied to loads or load effects and anchors, factors are applied to resistances rather than to
The collapse of the Ferrybridge cooling towers, noted above, material strengths, for two main reasons: (a) the designs often
illustrates the danger of combining loads too early in a involve load testing, which leads directly to resistances, and (b)
calculation. Equation 2.1 includes factors on both loads and because the construction process may change the ground
load effects. Eurocode 7 allows either to be used, but in either strength, a major uncertainty exists in deriving resistances from
case requires different factors on permanent and variable loads. the strength of the ground before construction; the resistance
In general, it could be difficult to distinguish the effects of two factors can be seen as factors on this process (or “model”) or on
different loads after they have been combined into a single load the strength of the ground at the interface with the pile or
effect. The situations shown in Figure 2.4 are again noted as anchor. There are some situations in which factoring loads at
examples. source leads to unreasonable situations, especially in the design
This issue is particularly critical to the design of spread of retaining structures. For these, EC7 allows the factors to be
foundations because calculated bearing capacity is very applied to the effects of the loads, and this is used where
sensitive to inclination and eccentricity of load. To augment appropriate in “Combination 1” of DA1.
information culled from literature, the author has discussed this In DA2 (Figure 2.9, partial factors are applied to loads and
with various experts. In the development of Japanese codes, to ground resistances. In a variant of DA2, DA2*, the
Shirato (2008) notes that factors are generally applied to loads equilibrium calculation is carried out using unfactored
before combination, though the client body for railway (“representative”) loads, and the factors are applied to derived
structures requires the use of unfactored loading, with all safety load effects. It has been found that DA2 and DA2* are
factors on resistances. For the Canadian Highway Bridge unsuitable for slope stability problems, so most countries which
Design Code, Green (2008) notes that loads are to be factored have adopted DA2 use DA3 for slope stability.
before being combined, so the inclination and eccentricity used In DA3, factors are applied to material strength and to loads
in the ULS calculation are based on the factored loads. Becker simultaneously, in contrast to the two-combination approach of
(2008) notes that this question has been debated in North DA1 in which they are applied to the two separately and the
America, without unanimous agreement. The AASHTO LRFD results compared. DA3 has been adopted mainly for slope
Bridge Design Specifications (2008) uses factored loads to stability problems only, though a few countries propose to use it
compute eccentricity and Figure C11.5.5-2 shows that earth for foundations and retaining walls, with factors quite different
pressure forces are factored before being combined for from those shown in Figure 2.9 in most cases.
calculation of eccentricity, stating “Abutments, piers and Simpson (2007) has emphasised that in many designs
retaining structures and their foundations and other supporting involving slopes, retaining structures and foundations ground
elements shall be proportioned for all applicable load and structures have to perform together and if a failure occurred
B. Simpson et al. / State of the art Report: Analysis and Design 2881

Figure 2.9: Factors proposed by CEN for the three Design Approaches
The author has found that the results of DA1 and DA2 are in
most cases fairly close, though DA2* may produce more
it would involve both of them The main advantages of DA1 economic, less safe designs in some cases. As noted above,
were stated to be: previous practice varies between countries. For the design
a. Consistent design of complete ground-structure systems problem shown in Figure 2.12, Orr(2005) found similar results
with the same two sets of factors. for DA1 and DA2. DA1 Combination 2 gives the design length
b. More consistent reliability across a wide range of for DA1, which is 4.73m, and the design maximum bending
geotechnical problems including foundations, slopes and moment is 163 kNm. The design length for DA2 is 4.69m,
retaining structures. which is marginally less than the DA1 length, while the
c. Factoring materials mainly at source, better control of safety maximum design bending moment is 177 kNm, which is
is obtained in frictional materials which give non-linear slightly greater than DA1. There is no application of DA2* in
relationships to resistances. this example, and DA3 is equivalent of DA1 Combination 2.
d. By factoring loads mainly at source, better control of safety It is hoped that some convergence of these approaches will be
in cases where loads tend to cancel each other. achieved in the future. The combined used of DA2 and DA3,
e. Ready use with finite element methods. adopted by some countries, is similar in many ways to the two
The main reason for adoption of DA2 has usually been stated combinations of DA1, suggesting some progress towards a
to be that it is more like previous practice and produces results common approach.
consistent with previous practice. This has been particularly
used in support of DA2*. 2.7 Some specific examples

2.7.1 Spread footings


For spread foundations with vertical loading, EC7 Design
Approaches 1 and 2 generally give similar results. For example,
the factors in Figure 2.9, Orr (2005) calculated widths of
foundations within 10% of one another for a typical example. It
is found, however, that the sizes of footings computed generally
imply high bearing pressures, even for the unfactored case, and
it is likely that serviceability checks will dominate the design,
leading to bigger footings. Footings designed to ASHHTO
(2008) strength limit state would be considerably larger than
required by the EC7 ULS. As seen in Figure 2.7, EC7 Design
Approach 1 follows the trend proposed in recent studies by
Foye et al (2006b), while Design Approach 2 allows for less
uncertainty of behaviour at high angles of shear strength.
For combinations of vertical and horizontal loads, however,
DA1, DA2 and DA2* may differ appreciably because the load
factors are applied at a different point in the process. Figure
2.13 shows an example published by Vogt and Schuppener
(2006) which provides an illustration of the effect of factoring
loads before or after they are combined. This involves the
calculation of the width of a strip footing subject to inclined
eccentric loading, based on ULS bearing capacity calculations.
A range of the characteristic value of the horizontal force
Figure 2.10: Adoption of the three Design Approaches by the various HQ,k is considered. Figure 2.14 shows that Combination 1 of
countries in Europe (after Bond & Harris 2008). DA1 requires two separate action combinations to be
considered, applying 1.5 to the variable horizontal action, and
2882 B. Simpson et al. / State of the art Report: Analysis and Design

Figure: 2.11: Factors adopted by the United Kingdom for Design Approach 1.
account of the separate factors on permanent and variable
load, or of the possibility that the permanent vertical load is
favourable.
The results of calculations of the required footing width for
varying ratio Hk/Vk are shown in Figure 2.16, using the bearing
capacity equations in Annex D of EC7. Results for DA1 and
DA2 are fairly similar, and DA3 is more conservative, as
expected. However, DA2* gives a markedly less conservative,
and so less safe design. In Figure 2.17, the factors of safety
available on the single quantities (a) γϕ on tanϕ′ and (b) γQ on
HQ have been calculated using the footing widths obtained by
DA2*. This shows that DA2* is equivalent to a single factor of
safety slightly greater than 1.2 on tanϕ′ or a variable factor on
HQ which falls to about 1.17 at HQ,k/VG,k of 0.4, the
approximate limit of this ratio for sliding (EC7 allows δ=ϕcv′ for
concrete case on the ground). As the only safety factor in the
calculation, a value of 1.17 on the variable load is remarkably
low; a factor of 1.5 is generally applied to variable loads..
Figure 2.12: Embedded retaining wall considered in the Dublin The proponents of DA2* argue that the results it obtains are
workshop of 2005 in line with previous experience. However, the author submits
that: (a) it is unlikely that there is a significant database of
footings which have, in practice, been loaded to the high H/V
ratios considered here (though it is very important that footings
which have to be designed for high ratios are reliable); (b) the
approach taken for bearing capacity is inconsistent with that
taken for sliding, which is likely to cause confusion; (c) the
approach taken for bearing capacity is inconsistent with that
taken for structural design, which is also confusing; (d) as noted
above, as the only safety factor in the calculation, a value of
1.17 on the variable load is remarkably low.
In practical use, design rules are often pushed to extremes
not anticipated by code drafters. For this problem, it is
desirable to check what happens if the lever arm of the
horizontal load is larger. It is assumed here that structural
Figure 2.13: Inclined, eccentric loading on a footing designers would factor the vertical and horizontal actions
independently. Figure 2.18, for a force at 10m above the base,
1.35 to the permanent vertical action if it is unfavourable (red shows that at a relatively modest ratio Hk/Vk= 0.13 the
dashed arrow), or 1.0 if it is favourable (red solid arrow). This eccentricity of the resultant load used in structural design
is consistent with EN1990 and the Eurocodes for structural exceeds the half width of the footing derived by DA2*,
design, and in principle it is also required for DA2 and DA3. implying that the resultant load passes outside the base of the
Figure 2.15 shows similar diagrams for Combination 2 of footing, which cannot give equilibrium. This will clearly cause
DA1 and for DA2 and DA2*. In DA2*, for consideration of consternation and confusion to the structural designer, and
bearing capacity, the horizontal and vertical loads are combined illustrates the inconsistency of DA2*.
into a single resultant before the load factors are applied, so the
inclination and eccentricity of the design action effect takes no
B. Simpson et al. / State of the art Report: Analysis and Design 2883

Figure 2.14: ULS design resultant actions derived for DA1.

Figure 2.18: Eccentricities in DA2 and structural design, for 10m lever
arm

Figure 2.15: ULS design resultant actions derived for DA1, DA2 and Figure 2.19: Eccentricities in DA1 and structural design, for 10m lever
DA2* arm

Figure 2.16: Footing widths calculated for ULS design

In contrast, Figure 2.19 shows that for DA1 the resultant Figure 2.20: Eccentricities in DA1 and structural design, for 100m lever
load always lies within the width of the footing; Combination 2 arm
is more critical than Combination 1, and so determines the
width. Figure 2.20 shows that DA1 even accommodates a 2.7.2 Piled foundations
much more extreme lever arm for the horizontal load of 100m. Traditional pile design has generally used overall factors of
This is guaranteed since the geotechnical design is checked for safety in the range 2 to 3, or even up to 4 in the case of tension
the same loading as the structure, consistently, in Combination loading. Often the lower end of this range is used when the
1, and for this extreme case it is Combination 1 which design is verified by load testing; and often larger values are
determines the footing width. For such an extreme case, much applied to base resistance than to shaft resistance. A major
of the width of the footing might be redundant and the structure reason for the use of relatively high factors has been that they
could be replaced by an A frame, but all the considerations ensure serviceability as well as safety against ultimate failure.
noted above would still apply. It is therefore necessary for developers of modern codes to
decide whether they will perpetuate this situation or try to
separate ULS from SLS. If a separation is attempted, it will not
be possible, in many cases, to calibrate factors needed for ULS
against existing practice, and SLS will often dominate the
design, at least for larger diameter piles.
EC7 requires a “characteristic” pile resistance which is a
“cautious estimate”. In practice this can be derived from
calculation based on ground testing or from load testing of piles,
or a combination of the two. Indeed, EC7 specifically requires
that calculation methods must be verified by testing, and test
results must be checked by calculation {7.4.1(1)}, but it does
not clarify how information from these two sources can be
combined when both are used together. It also states that “For
2.17: Total factors of safety implied by DA2 on tan ′ and HQ, taken piles bearing in medium-to-dense soils and for tension piles, the
singly.
2884 B. Simpson et al. / State of the art Report: Analysis and Design

safety requirements for the ultimate limit state design are or ground strength, such as the use of piles or anchors used to
normally sufficient to prevent a serviceability limit state in the hold down the slab as shown in Figure 2.22(b).
supported structure” {7.6.4.1(2)}. Thus for piles, in contrast to
other design elements, the code implicitly requires that factors
for ULS design be chosen so that SLS will be satisfied. Hence
the overall factors of safety adopted should probably be similar
to previous practice.
Several European countries have reported that the partial
factors proposed by EC7 imply overall factors of safety much
lower than they would traditionally have used. The code is
specific about the process of deriving design resistances from
load tests, applying “correlation factors” ξ to the mean and
lowest test results. Much less detail is given about calculation
of resistances from ground strength tests, however. This means
that each country has some freedom to specify in rather more
detail how pile resistances will be calculated, and the values of
partial factors may also be varied nationally, as for other design (a)
elements.
Bond and Simpson (2009) describe the background to the
derivation of the factor values required by the UK National
Annex to EC7 (BSI 2007). The partial factors γ have been
increased compared with the values in the base version of
EN1997-1 (see Figure 2.9); the correlation factors ξ have also
been increased and additional “model factors” have been
imposed in the derivation of characteristic shaft and base
resistances from characteristic ground strengths. Both the
model factors and partial factors γ have been varied according
to the extent of load testing which supports the calculations.
The overall effect for ULS calculation is shown to be consistent
with previous design, and the characteristic values of resistance
(b)
are suitable for serviceability calculations as, for example, in
settlement reducing piles which might be loaded to their
ultimate capacity in the working state. Example calculations are
provided by Bond and Simpson (2009).

2.7.3 Water pressures


Application of factors to water pressure is fraught with
difficulty and can easily lead to unreasonable situations. Orr
(2005) reported calculations for the situation of potential
hydraulic heave shown in Figure 2.21. He found that the
calculated allowable height of water H could vary from 2.78m
to 6.84m due to application of the same factors, taken from
EC7, at different points in the calculation. (c)
Figure 2.22: (a, b) ‘UPL’ and (c) ‘HYD’ situations in EC7.

Modern codes tend to provide a single factor for “dead


loads” or “permanent actions”. It is arguable that static water
pressure in the ground is a permanent action. In some cases,
design can proceed, apparently reasonably, with a factor (eg
1.35) applied to water pressure, but in other cases applying a
factor to water pressure is unreasonable. For example, Figure
2.23 shows a deep basement or shaft in ground with a high
water table; increasing water pressure in the ground near the
base of the shaft by a significant factor (eg 1.35) would
represent a physically impossible situation and would constitute
Figure 2.21: Hydraulic problem considered in the Dublin workshop of an unreasonably onerous design case.
2005 The AASHTO code {10.6.3.1.1} requires that “bearing
resistance shall be determined based on the highest anticipated
In design of structures in the ground, water pressure may position of groundwater level at the footing location”, but it
constitute an additional load to be considered along with others, does not apply factors to water pressures for foundation or
to be accommodated by strength requirements in structure retaining wall design, despite factoring effective earth pressures.
and/or ground. In addition to this, two particular situations can This appears to imply that in a situation where ground water
be identified in which water pressures are principally balanced pressure is dominant the design would rely almost entirely on
by other loads (weight of structures or ground): uplift failure the resistance factors in both the ground and the structure.
and hydraulic failure, termed UPL and HYD in EC7, as Eurocode 7 requires the designer to consider two situations.
illustrated in Figure 2.22. EC7 provides factors of safety to be For ULS, the design should use “the most unfavourable values
used in checking these limit states, but it is not clear about that could occur during the design lifetime of the structure”,
where in the calculation they should be applied, leading to the whereas for SLS “the most unfavourable values which could
confusion noted by Orr. In practice, these states are often occur in normal circumstances”. It also allows that design
complicated by the involvement of some element of structural
B. Simpson et al. / State of the art Report: Analysis and Design 2885

water pressures for ULS could be derived by applying a partial 1. Apply DA1-1 using reasonably cautious water levels (most
factor to characteristic water pressure or by raising the water unfavourable in normal circumstances) and apply 1.35 to
level. Figure 2.24 compares water pressures for a 10m high structural load effects - bending moments etc.
wall with a 2m difference between these two levels. Starting Alternatively, in some cases, it is necessary to apply this
from the “characteristic” water pressures (line A), it can be seen factor to all permanent actions, including water pressures.
that merely applying a factor (obtaining line B) may Particular caution is needed if non-linear behaviour of
underestimate the effect of raising a water level (line C); this is structures is expected.
particularly the case when thrusts or bending moments are 2. Apply DA1-2 using worst credible water levels (most
considered. On the other hand, factoring water pressures which unfavourable in design lifetime), unfactored.
are already “the most unfavourable values that could occur
during the design lifetime of the structure” (obtaining lines D or 2.7.4 EQU
E) seems unreasonable from both probabilistic and physical Situations in which independent loads tend to cancel each
points of view. other’s effects were discussed above when considering working
state design. In a partial factor approach, the design is generally
made robust by applying different factors to the two loads. A
further problem is sometimes encountered when two loads
which are not independent, but arguably two parts of the same
load, have a cancelling effect.
Eurocode “Basis of design”, EN1990, considers that such
loads come from a “single source”, and notes: “For example,
all actions originating from the self weight of the structure may
be considered as coming from one source; this also applies if
different materials are involved.” When there is good reason to
believe that the two loads must occur together, it seems
unreasonably severe to factor them separately. Nevertheless,
some allowance for variability of loads also appears necessary.
EN1990 terms this situation the “EQU limit state” and requires
a check by applying small load factors, generally 1.1 and 0.9, to
the two parts of the single source load.
Schuppener et al (2009) have debated the significance of
EQU to geotechnical design, presenting several alternative
Figure 2.23: Deep basement or shaft in ground with a high water table
approaches without an agreed conclusion. In Figure 2.4(a), the
approach to this issue determines the design bending moment in
the column and changes the design loading in the piles, possibly
putting one of them into tension; in Figure 2.4(b) it affects the
design load in the anchor. The author’s opinion is that “the
whole system, both structure and geotechnics, should be
designed for all the relevant design situations or load cases”;
alternative approaches are inevitably more confusing and run
the risk of leaving gaps between design cases, in which either a
critical situation is not accommodated or design appears
impossible.

2.8 Economy and sustainability


Figure 2.24: Water pressures for a 10m high wall Codes influence both the safety of constructions, probably their
primary aim, and their economy of constructions. With modern
The UK National Annex to EC7 (BSI 2007) points designers appreciation of the need to reduce usage of materials and
away from simple factoring of water pressures, though it leaves energy, it is clearly desirable to produce more economic
the option open. It says that standard load factors “might not be designs, which might be assisted by reductions in values of
appropriate for self-weight of water, ground-water pressure and partial factors, where possible. It is commonly stated that
other actions dependent on the level of water”, and “the design geotechnical failures, ultimate or serviceability, are rarely
value of such actions may be directly assessed in accordance caused by having factors of safety slightly too low. It is also
with” the EC7 principles noted above, or alternatively “a safety reported in Eastern European countries that the factors proposed
margin may be applied to the characteristic water level”. in EC7 (as published by CEN) lead to designs which are
In contrast, Bond and Harris (2008) provide two arguments uneconomic compared to their previous practice, which they
in favour of factoring water pressures when designing retaining found satisfactory. This suggests that the proposed factors
structures: (a) structural engineers have traditionally applied might be reduced, and so a proposal has been put to CEN to
partial factors to retained liquid loads and ground water investigate further the designs used in Eastern Europe, with the
pressures, and (b) it is common for geotechnical engineers to aim of reducing the proposed factors. If it is possible, it will be
perform calculations using unfactored parameters, including necessary to compare records for safety and durability between
water pressures, and then to apply a factor of safety to the societies adopting different designs. Hopefully, common
resultant structural effects, giving the same result as factoring international codes will provide a framework in which this
both the effective earth pressures and water pressures. might be achieved.
Provisions for safety in relation to water pressure, both free Individual factors must be reduced with care since factors on
water and ground water, remain under debate within the EC7 one variable contribute to the overall robustness of the design,
Maintenance Group. Given such uncertainty, making more than perhaps giving unanticipated benefits . For example, the public
one check, effectively a parametric study, may be advisable. enquiry into the Nicoll Highway collapse in Singapore (Magnus
For EC7 Design Approach 1, the author has found that for most et al, 2005) noted that one of the contributory causes was factors
situations the following strategy gives reasonable results, of safety lower than code requirements. While this was not the
though critical review is always necessary: prime cause of the failure, Simpson et al (2008) have argued
2886 B. Simpson et al. / State of the art Report: Analysis and Design

that use of standard factors would probably have prevented test data) is not presented herein, it clearly remains a valid
collapse, mitigating the effects of more serious deficiencies. approach provided measured pile resistances can be justified by
Further more, while reduced factors might not lead to collapse, calculation methods derived in keeping with Method 2.
they could result in more durability problems in later years. It is
therefore very important that large scale surveys of performance 3.2 “Mega” piles: an introduction to vertical design
are undertaken as an aid to introducing reductions in code
factors. Where appropriate, large diameter piles are used in onshore
Both economy and safety should be aided by more realistic construction for infrastructure projects or high-rise building
evaluation of the necessary resistance of structures and ground. development both of which generate high loads. The upper
To this end, the Japanese development of Performance Based limit of such piles is typically 1.5m in diameter, occasionally
Design aims to provide a framework which incorporates limit 2.1m or 2.4m in diameter, due to limitations on the size of
state design and is particularly applicable to seismic situations. construction plant. In offshore locations “mega” piles with
Probably the biggest influence geotechnical codes can have diameters in excess of 5m are now possible owing to larger
on both safety and economy, however, is to emphasise the construction plant and the economics of construction (speed of
importance of good geotechnical process, desk study, ground construction is critical to economy off-shore). The design of
investigation and choice of construction type. Calculations and such large diameter piles is presented below with a bias towards
factors, necessary for analysis and design, are important, but developments in offshore design practice.
secondary. To quote EC7: “It should be considered that Traditional methods of pile design rely heavily on
knowledge of the ground conditions depends on the extent and observation leading to empirical equations for assessment of
quality of the geotechnical investigations. Such knowledge and axial (geotechnical) capacity. This is still true, even though, as
the control of workmanship are usually more significant to pointed out by Randolph (2003a), the scientific understanding
fulfilling the fundamental requirements than is precision in the of soil behaviour in general and around piles is continually
calculation models and partial factors.” improving. Such improvements in understanding allow
observed pile behaviour to be better explained, thereby allowing
for better design. Design based on empiricism is nevertheless
3 DEEP FOUNDATIONS the norm for assessment of pile capacity under axial loading,
although theoretically based equations are sometime used for
assessment of base resistance in some instances.
3.1 Introduction This general summary of the current state of design is
illustrated by the much used API (2000) “Recommended
The 2005 ICSMGE Osaka conference included a State-of-the- Practice for Planning, Designing and Constructing Fixed
Art paper by Mandolini et al (2005) titled “Pile foundations: Offshore Platforms” document where much of design guidance
experimental investigation, analysis and design” and a further relies on empirically assessed resistance similar to Equations
paper by Randolph et al (2005) titled “Challenges of Offshore 3.2 and 3.3 below. (It is understood that a revised version of the
Geotechnical Engineering”. In Mandolini et al (2005), the API document will be released in due course which will be
authors reviewed both single pile behaviour and pile group updated to included much of the work referenced below for
behaviour, considering both experimental results and analysis. driven piles offshore.) The following review of pile design
The paper by Mandolini and his co-authors remains a thorough shows that while empiricism remains at the forefront of pile
review of these aspects of pile design and analysis. In Randolph design, soil mechanics theory is continually developing and
et al (2005) a summary of offshore ground investigation is providing important insights into the assessment of pile capacity
presented along with design and construction aspects of piled and behaviour. Soil mechanics should be seen as a route to
foundations and spread footings (suction caissons, gravity improving our understanding of pile behaviour and, with
structures etc). Four years further on, the ISSMGE committee recourse to best practice design methods, to informing ground
for the Alexandria conference has asked for this short investigation techniques.
contribution on deep foundations as part of a more general In the following, care is taken to present equations and data
paper on design and analysis; the committee suggested that for ultimate conditions. Where “allowable” or working
large diameter or “mega” piles could be discussed. This capacities are quoted, these are clearly identified. When
contribution provides an update on design processes for single choosing the base soil parameters (e.g. characteristic values in
piles in soil and rocks using design processes from both the Eurocode design) the designer must accommodate the basis
offshore and onshore developments with bias towards large of the code adopted. For example, an average value of shear
diameter piles and vertical loading. strength may be appropriate to one code while in another the
In addition to the sections on load-carrying capacity, parameter may need to be a cautious estimate. Failure to
comments on sustainable aspects of pile design are provided in understand the basis of the code will result in inappropriate and
view of the recent progress in this area and the lack of inclusion potentially unsafe designs.
in previous ICSMFE state-of-the-art reports. In assessment of the ultimate limit state (ULS) many
Prior to presenting a limited number of pile design methods conventional design processes protected the serviceable limit
it is worthwhile setting the scene for acceptable design. state (SLS) for conventional structures by limiting allowable
Eurocode 7: Geotechnical design - Part 1: General rules (BSI, working load to a percentage of the skin friction (e.g. skin
2004) provides reasonable requirements for acceptable design: friction / 1.2, in addition to a factor of safety on total capacity)
The design [of piles] shall be based on one of the following or by ensuring mobilisation of base resistance is small. As
approaches: codes become more sophisticated and as construction extends
• [Method 1] the results of static load tests, which have been beyond historic limitations (e.g. large diameter piles), there is a
greater need to assess pile settlement as part of the design
demonstrated, by means of calculations or otherwise, to be
process with attention paid independently to the rate of
consistent with other relevant experience;
mobilisation of shaft and base resistances. Clearly where there
• [Method 2] empirical or analytical calculation methods is direct comparable experience (loading, foundation size and
whose validity has been demonstrated by static load tests in ground conditions) the requirement for explicit assessment of
comparable situations. settlement is reduced. Where such comparable experience is
Within this report attention is paid to the second of these two absent then explicit assessment of settlement should be the
methods where pile load test data is used to develop design norm.
methods. While the first method (design by site specific pile
B. Simpson et al. / State of the art Report: Analysis and Design 2887

As a final introductory comment, no discussion on the • Where pile size is larger than test pile diameters, then
structural capacity or durability of piles is provided herein. interpretation of test data must allow for the effect of
Such considerations are clearly required for a design in practice diameter on base and shaft capacities and overall settlement
and should accommodate applied actions and construction to derive ULS and SLS conditions. As noted in sections
tolerances as well as changes to the structure or dimension of 3.2.2 and 3.2.5 below the measured pile shaft resistance
the pile with time (e.g. corrosion of steel piles). (stress) varies as pile diameter increases thereby not
allowing linear extrapolation from small to large diameter
3.2.1 Pile design based on offshore experience: soil piles when assessing shaft capacity.
The following paragraphs present a summary of offshore pile • Finally, where piles are constructed in new strata, the
design for silica sands and clays. Comment is made on pile behaviour of the strata during pile construction and
capacity in carbonate soils in a following section. subsequent loading must be understood.
Conventional design of piles usually adopts an equation as As an example of the above comment on soil strata, pile base
follows: capacity calculated using a constant value of ϕ' and a
recognised bearing capacity factor (Nq ) such as Berezantzev et
Q = (Qb + Qs)= (Abqb + •A
Asqs) (3.1) al (1961) will fail to recognise the subtleties of soil behaviour at
varying stress levels as illustrated by Bolton (1987), Fleming
Where Q is pile resistance and q is limiting or failure stress, (1992) and as extended by Cheng (2004). In the context of pile
A is area and subscripts b and s are base and shaft respectively. base bearing capacity, higher levels of confining stress result in
The method for calculation of qb and qs has, in the past, been a reduction of the ability of the soil to dilate. This prevents the
relatively simple and commonly took the form of: peak angles of friction seen on samples tested at low effective
stress being mobilised. This observation is valid for siliceous
In sand: sands but becomes critical when softer soil grains are involved
due to early onset of particle crushing and further suppression of
Q = Abqb + •A
2
Asqs = ¼ D Nq 'v + D •(K 'v tan )dL (3.2) dilatancy (Klotz and Coop 2001).

and in clay: 3.2.2 Offshore: Driven piles – a CPT approach


The environment in which offshore piles are constructed
Q = Abqb + Asqs =
2
¼ D Nc su base + D •(( su L)dL (3.3) combined with the requirement to carry large loads has resulted
in offshore piles being significantly larger than their onshore
where the terms are as follows: counterparts. Driven tubular piles installed offshore can be 5m
K earth pressure coefficient on the pile shaft and larger in diameter, dwarfing typical construction onshore.
angle of interface friction (often expressed as a fraction of The viability of such large piles is linked to progress in pile
the shear strength, e.g. 2⁄3 ϕ) installation technology and to the research into axial pile
shaft adhesion factor applied to su behaviour resulting in new design methodologies as referenced
D pile diameter below. Whilst these new design methods are driven by the
L length for assessment of pile shaft friction requirements of offshore construction, they can equally well be
su undrained shear strength of the soil (clay). Subscripts used to inform onshore pile design when similar construction
“base” and “L” illustrate the likely different values in su that techniques are used and ground investigations carried out. They
will be used in the calculation at the base and incrementally may also be used as starting points for development of methods
along the shaft. appropriate to onshore practice. The design methods typically
These equations are populated with a combination of use (sea-bed driven) CPTs to provide a profile of soil resistance
theoretical (Nq, Nc) and empirical (K, Į) parameters. When together with laboratory testing to provide further parameters
using such equations there is significant guidance on the choice (density, interface strengths, magnitude and variation of Ko
of the design constants and limiting stresses for situations etc.). Where CPT data is not continuous appropriate averaging
within bounds of comparable experience. Such guidance has techniques must be used to generate a continuous profile.
been distilled into books such as Poulos and Davis (1980) and The UWA-05 (University of Western Australia 05, Lehane
Fleming et al (2008) among others. Whilst these equations and et al., 2005) method is used to illustrate the developments in
their empirical parameters have come about from experience pile design in siliceous sand, albeit, as noted by Monzon (2006),
and remain of practical importance, they do not provide alternatives exist in the form of ICP-05 (Jardine et al) and NGI-
adequate confidence for situations where there is no comparable 05 (Clausen et al, 2005). The UWA report deals with piles
experience. Such situations include new construction installed (driven) in silica sands only. All methods mentioned
techniques, sites with no previous pile performance data, above differentiate between closed and open ended driven piles
changes in pile dimensions compared to test data and where pile owing to the different strain fields which develop during
construction is in new strata. Taking these conditions in turn: driving.
• Where new construction techniques are adopted there is Base resistance for close ended driven piles in sand is related
little option but to carry out high quality instrumented load to CPT cone resistance qc as follows:
tests to investigate pile behaviour. The results of such tests
should be interpreted so as to distinguish between pile shaft Qb = Ab qb0.1c (3.4)
and base behaviour (load settlement response). The results
should also be assessed against the best available qb0.1c = 0.6qc ave (for jacked piles the factor 0.6 can be increased
calculations in order to have confidence of the validity of to 0.9)
test data. Added confidence is gained as more piles are qc ave the average value of qc over a distance 1.5 times the pile
tested, a process formalised by Eurocode 7 Part 1 (BSI, diameter above and below the pile tip (further reference
2004). is provided in Xu and Lehane, 2005).
• Where a new location is being developed, an understanding
of the soil is critical with, as a minimum, assessment of The subscript 0.1c denotes the mobilised resistance at a
geomorphological and geological situations (ground model), settlement of 10% (0.1) of the base diameter for a closed ended
mineralogy and soil consistency (relative density, strength (c) driven pile; the equation is clearly empirical.
and stiffness data) being required to correlate to other
locations.
2888 B. Simpson et al. / State of the art Report: Analysis and Design

For open ended piles experience (Lehane and Randolph,


2002) shows that in siliceous sand, piles loaded under static
conditions tend to plug when the height of plug inside the pile at
the end of driving is greater than 5 times the pile internal
diameter (5Di). This observation was made on piles of up to
1.5m diameter, care must be exercised for piles with greater
diameters and additional research is required to allow general
adoption of this observation. The result of this is that shaft
friction is generally linked to the external face of the pile only
and base capacity is modified as follows:

Qb = Ab qb0.1o (3.5)

where
Figure 3.1 Relationship between Effective Diameter, pile internal
qb 0.1o ª
2
§ D *· º (3.6) diameter and pile external diameter.
= «0.15 + 0.45¨ ¸ »
qcave ¬« © D ¹ ¼»

The subscript “0.1o” denotes the mobilised resistance at a


settlement of 10% (0.1) of the base diameter for an open ended
(o) driven pile. While it is assumed that the pile plugs, the open
ended pile “capacity” is less than that of a closed ended pile due
to the higher stiffness exhibited by the closed ended piles
compared to an open ended pile. This observation in stiffness is
a common thread seen in other empirical based design methods
as in Section 3.2.3 below. It is also consistent with previous
comments in Section 2.6.2 where partial factors for ULS pile
design generally provide a degree of comfort for SLS
considerations. D* is the effective diameter of the pile’s base
and is defined as (all pile diameter dimensions in metres): Figure 3.2 Relationship between end bearing stress on full pile
diameter and average cone resistance at pile tip with pile diameter as a
D * §¨ ­° ª D º 0.2 ½°§ D 2 · ·
0.5
(3.7) function of Di/D.
= 1 − min ®1, « i » ¾¨¨ i2 ¸¸ ¸
D ¨© °̄ ¬1.5 ¼ °¿© D ¹ ¸¹ piles. Measured pile shaft friction is linked with cone resistance
and the fatigue induced at the interface between pile and sand
For open ended piles qc ave is the average value of qc over a by pile driving as well as the interface friction (sand to pile) and
dimension of 1.5 times D* above and below the pile tip and not radial effective stress acting on the piles shaft due to the
the external diameter, D, as for close ended piles. In Equation displacement nature of the pile. The empirical correlation
3.7 the “min” value is referred to as the “final filling ratio” between pile shaft friction qs and cone resistance qc is expressed
FFR) and is, in the absence of direct measurement of the plug in terms apparently dissimilar to the shaft component in
length inside the pile, an approximation allowing for the plug Equation 3.2 above. The term Kı'v ( = ı'h ) in Equation 3.2 is
length. It should be noted that this FFR approximation only replaced by a function of qc which accounts for horizontal
holds good for situations where the pile tip has penetrated to at effective stress working on the pile shaft. The horizontal
least 5 times the internal diameter of the pile into the sand layer. effective stress is a result of the post-installation conditions at
Where this is not the case, assessment of non-plugged pile the pile face combined with the dilation of the soil adjacent to
capacity will be necessary. the pile which is mobilised during pile loading. The latter term
Value of D*/D are presented in Figure 3.1 for Di /D in the can be compared to the enhanced resistance derived in a
range of 0.98 down to 0.0. Di /D = 0 for a closed ended pile and constant normal stiffness shear box test, where dilation results
0.98 for a thin wall pile. The limiting conditions occur when in an increase in confining stress and therefore shear resistance.
the internal diameter equals 1.5m as can be seen in Equation 3.7 For large diameter piles the component of shaft friction derived
above. Figure 3.2 shows the variation in the ratio of end from dilatancy of the soil at the pile interface is relative small
bearing stress (average over the full cross sectional area) and allowing a simpler expression for shaft friction to be presented
cone resistance in the vicinity of the pile tip with pile diameter (Lehane et al, 2005b):
for a range of Di/D values; for closed ended piles the ratio is
0.6, falling to less than 0.18 for relatively thin walled tubes. ª § h ·º
−0.5
(3.8)
× tan δ cv
0.3
As would be expected, for a zero dimension of the internal qs = a.qc Ars × «max¨ 2, ¸»
¬ © D ¹¼
diameter Di (i.e. a closed ended pile), the value of Qb is the same
as a closed ended value in Equation 3.4 above.
Assessment of pile shaft capacity for open and closed ended
qs = 'h x Fatigue factor x Coefficient of friction
piles has been combined into a single method in siliceous sands
a a constant equal to 0.03 for compression piles and 0.0225
(Lehane et al, 2005b). This is possible due to the tendency of
for tension piles.
open ended piles to plug during static loading thus negating the
Ars the effective area ratio for the shaft.
need to assess skin friction on the inside of the piles tube..
h distance of the section of pile shaft above the pile tip
While the method of assessment of pile shaft capacity is the
(factor allows for fatigue during installation. For long
same for open end closed ended piles, closed ended piles
piles the top section of the pile has large value of “h”
mobilise high resistances due to higher radial stresses acting on
resulting in reduced skin fiction at that location compared
the pile shaft resulting from the displacement nature of these
to locations further down the piles (assuming similar
values of qc).
B. Simpson et al. / State of the art Report: Analysis and Design 2889

cv
constant volume angle of interface friction (sand on steel), cast-in-place rather then driven. Whilst onshore and offshore
correlated to the mean particle size (approx = 23° for D50 pile sizes and loading conditions are different, the requirement
= 1mm rising to an upper bound of 28.8° for D50 = for overall economy is comparable albeit the manner in which
0.2mm). economy is achieved is not necessarily the same. For offshore
construction the cost of plant (and thereby time on site) is a very
Ars is equal to 1.0 for closed ended piles and can be important factor whilst for onshore conditions material usage
calculated for open ended piles from the following: Ars = 1- has a larger impact on overall cost.
IFRmean (Di / D)² where IFRmean is the incremental filling ratio When considering the technical aspects, rather than the
and, in the absence of measured data is equal to IFRmean = economic aspects of design, the soil around onshore piles
0.2 0.3
min(1,(Di /1.5) ). Values of Ars are presented in Figure 3.3 experiences disturbance in a similar manner to that of offshore
where it can be seen that a thin walled pile (Di/ D = 0.98) has piles for similar pile types (bored, driven etc). Installation
approximately 40% of the shaft friction of a closed ended piles processes modify in-situ stresses, and possibly ground strengths
(Di / D = 0) due to remoulding and shearing, to such an extent that soil
mechanics alone is again rarely able to predict pile performance
accurately. Hence, in the geotechnical design of onshore piles
constructed in soil, empirical procedures dominate. An
illustrative example of this is well documented in French design
practice, now enshrined in Eurocode 7 Part 2 (BSI, 2007a),
using the Ménard Pressuremeter limit pressure to assess both
limiting shaft friction and base resistance for all main pile types
in a wide variety of soil types. Such design is useful as it is
based on a large set of pile observations and is appropriately
conservative. It is does not, however, lead to an understanding
of pile behaviour.
Published design guidance (BSI 2007a) can be supplemented
by site specific correlations to optimise pile design and to
accommodate different methods of pile installation. An
example of this is bored-base-grouted piles in dense sand
Figure 3.3 relationship between Effective Area Ratio (^0.3) and
diameter for various Di/D values. (Thanet Sand) used in East London (UK) for tall buildings and
occasionally in central London where deep infrastructure
It is noted that Equation 3.8 is appropriate for use with prevents founding at shallow depth (Yeow et al, 2005). Early
driven piles. Where piles are jacked then a similar approach can design procedures were based on pile load tests (Troughton and
be taken except that the fatigue factor (max (2,h/D)-0.5) must be Platis, 1989, Chapman et al, 1999). Pile base capacity was
modified. The approach presented by Jardine et al (2005) correlated with mean effective stress rather than the more usual
provides more data. vertical effective stress. Variations in the “apparent” relative
In order to allow for dilatancy the reader is referred to the density of the sand with depth (corrected SPT blow count
original paper to allow economic design of small diameter piles. reducing with depth) along with increasing fines content with
As noted above, the dilatancy effect does not increase the shaft depth could not, with confidence, be used to assess pile base
resistance significantly (more than 10%) for all but small capacity variation with depth. This resulted in the need for a
diameter piles (less than about 0.6m external diameter). more robust method of profiling pile end bearing resistance with
The above method of pile design is based on the results of depth.
load tests on a large number of relatively small diameter piles, The Menard Pressuremeter approach (Nicholson et al 2002)
as is fully acknowledged by the authors of the pile design was adopted and equipment extended to allow testing to 10MPa
method. The potential benefit of carrying out maintained load to allow the engineering implications of the changes in the sand
prototype scale pile tests is worth investigation as suggested by with depth to be accommodated in design (typical equipment is
White and Bolton (2005), and should be pursued where possible limited to 5MPa or less and would not allow direct limit
(e.g. for large off-shore wind farms where the benefits of scale pressure measurement). The accepted equation for pile base
may justify the initial cost of testing). capacity using the Ménard Pressuremeter limit pressure is:
While the above suggests that design development has
progressed significantly in recent years it is also noted that qb = ks (Plim - )+
h0 h0
(3.9)
significant research continues. This is well illustrated by a
recent research project led by the Norwegian Geotechnical ks empirical factor
Institute on pile testing and design. The project aims to test Plim limit pressure
relatively small diameter (0.3 to 0.4m diameter) driven open h0
horizontal total stress
tube piles in up to 6 varying soil types (see comments on the
need to test large diameter piles above). At each site, piles will The value of ks is purely empirical as illustrated by the
be installed and load tested over a period of 2 years to identify values of 1.6 for bored piles and 3.2 for driven piles in dense
how pile capacity increases with time from construction to two sands (BSI, 2007). This derives from differences in pile base
years post construction. Investigation of how loading affects stiffness for loading to 10% of pile base diameter (driven piles
pile capacity will also be carried out. While increase of pile are stiffer then bored piles in the sandy soils). For bored-base-
capacity with time is acknowledged, the research aims to grouted piles based in Thanet Sand, a value of ks of 2.5 was
provide greater clarity on the issue. Since pile testing is usually found to correlate with load test results. The value of 2.5 lies
carried out shortly after construction, design methods typically between 1.6 (conventionally bored) and 3.2 (driven) as
predict pile capacity at 10 to 20 days after construction. suggested by Nicholson et al (2002). The value is clearly linked
to the nature of base grouting and as with all empirical factors is
3.2.3 Onshore: Bored piles - a pressuremeter approach very much method related and cannot be taken as universally
Compared to offshore development, for a majority of onshore applicable.
developments there will be increased constraints in terms of As with all empirically based design methods the parameter
noise and vibration. The result of this is that, for urban (limit pressure in this case) is only valid if the conditions
pertaining at the time of ground investigation test are similar to
development at least, large diameter piles tend to be bored and
those governing the design of the foundation. If the conditions
2890 B. Simpson et al. / State of the art Report: Analysis and Design

(notably effective stress regime in the soil) vary between the revenue). The impact of such soil-structure interaction is that
time of the ground investigation testing and the design drilled and grouted piles (or possibly driven and grouted piles)
condition, these changes must be accommodated in the design. are a more suitable foundation type for such soils. Design
Such changes result from basement excavations or changes in recommendations, such as those of Randolph et al (2005)
groundwater table. Whilst the design approach is clearly suggest a peak shaft friction (qs (peak)) as per Equation 3.11:
empirical the method of correction must be based on something
more substantive in order to predict pile capacity at the design qs (peak) • qc (0.02 + 0.2 e
-0.04 qc/pa
) 3.11
condition. For the case of piles in sand, the method by
Troughton and Platis (1989) corrected base resistance by the where pa is atmospheric pressure (100kPa),
ratio of mean effective stresses ( m') at the design condition Similar behaviour is also well documented for piles driven
compared to the testing condition based on measured pile into chalk (low and medium density chalk) where the limiting
performance. Using this approach then the corrected pile base design shaft friction for driven piles is often 20 to 30% of the
capacity is: equivalent cast-in-place pile. Again this is due to the
contractant nature of the remoulded chalk and the ability of the
qb = ( m'design / m'GI) (ks (Plim GI - h0 GI
)+ h0 GI
) (3.10) chalk mass to arch around the pile shaft (Lord et al, 2002).
For initial assessment of pile behaviour in chalk or carbonate
soils, the reader is directed to Lord et al (2002) and Randolph
where subscripts “design” and “GI” indicate the working (2005) respectively.
condition of the pile and the conditions existing at the time of The above example highlights the need to obtain a thorough
the measurement of limit pressure. understanding of the effects of soil mineralogy and in-situ state
Menard Pressuremeter limit pressure can similarly be used to when carrying out pile design. For designs being carried out at
predict shaft capacity as described in BSI (2007a) where new locations and where the soil has not supported piles or
empirically derived graphs are presented plotting limit pressure indeed the type of pile being proposed, design must not rely
against shaft friction as a function of soil type and density. only on empirical rules developed elsewhere for different
Where a reduction in effective stress occurs between conditions. An open mind must be maintained and the ability to
pressuremeter testing and the design condition then it is load test piles must be incorporated into the programme and
possible, for conventionally bored piles in sands at least, to cost plan.
correct the limit pressure as a function of changes in horizontal
effective stress (§ įplim = plim test 8 įıh' / ıh'). Assessment of 3.2.5 Drilled and grouted piles – rock:
the initial ıh' requires a knowledge of stress history (Mayne and Drilled and grouted piles are commonly used in near-shore
Kulhawy, 1982) or by direct measurement. Change in ıh' (įıh') locations where rock exists below the seabed and the
may be made assuming įıh' = (Ȟғ' / 1-Ȟғ') 8 įıv', where Ȟғ' is the construction of direct bearing foundations (gravity structures) is
drained Poisson’s ratio (Troughton and Platis, 1989). A check not cost effective (e.g. where there are significant soft sediments
on the final ratio of ıh' / įıv' is necessary such that ıh' ” 1/ Ko NC above rock head). These foundations are common for port
8 ıv' where Ko NC is the normally consolidate value of Ko structures, bridge foundations and wind turbines as well as for
(§1−sinϕ') unless pile testing shows higher limiting values (such
oil and gas industry related structures. Such foundations have
as ıh' ” Kp 8 ıv'). It would be reasonable to assume that such
become more common in recent years as a result of
an approach is appropriate only to a reduction in limit pressure
and not an increase. Such changes in effective stress could be developments in piling technology allowing single piles to carry
due to wide excavations or changes in groundwater level large loads (often limited by the structural capacity of the pile
following completion of the ground investigation. shaft). These positive developments have also coincided with
While pile design is dominated by empiricism our heightened awareness of the dangers of construction utilising
understanding of soil behaviour is more firmly rooted in science divers or working under pressure, as in the case of sunk
and theory. This is ably demonstrated by Ventouras and Coop caissons, further championing the case for drilled and grouted
(2009) who investigated the case of reducing pile base capacity pile foundations. In parallel with advances in construction
with depth by means of triaxial testing. The tests were carried capability there have also been consistent advances in design, as
out at stress levels appropriate to a loaded pile bases in order to presented below.
investigate the how the variations in soil grading and particle Shaft Friction: Classical design of rock-socket piles is
mineralogy could effect pile capacity. The results showed that empirically based (Rowe and Armitage, 1987, amongst others)
the increase in fines content with depth was the dominant factor and relates the unconfined compression strength of the intact
resulting in reduced pile base capacity with depth due to rock to the ultimate shaft friction (qs ) as follows:
reduction in the ability of the soil to dilate during shearing.
While the conclusions of the research are not such that pile qs = q ci
(3.12)
capacity can be predicted for conventional design, they do
provide better understanding on how piles behave. Such or
understanding results in better and safer design.
qs = c ci
(3.13)
3.2.4 Piles – Carbonate soils and chalk:
The above description of offshore driven piles and onshore qs unit shaft friction (MPa)
bored piles is related to silica dominated sands. Where q
, c
and are empirical constants
carbonate sands are present (often lightly cemented) experience ci
intact rock unconfined compression strength (UCS in
shows that bored piles (drilled and grouted piles) may be MPa)
preferable to driven piles. This is due to the likelihood of low
confining stresses acting on the shafts of driven piles caused by Values of c and are typically 0.2 to 0.4 and 0.5 respectively
the local collapse of the adjacent soil (installation damage) and (Hovarth et al, 1980). The range in the value of c is related to
the ability, thereafter, of the soil to arch around the pile. Such the roughness of the rock socket.
behaviour is well documented from the Rankin field of Western The parameter Įq is shown in Figure 3.4 and is obtained by
Australia (King and Lodge, 1988) where the cost of remedying combining Equations 3.12 and 3.13, giving:
failed driven piles came to A$300m at the time (at current –1
prices about US$500m excluding costs associated with loss of Įq = c ci
(3.14)
B. Simpson et al. / State of the art Report: Analysis and Design 2891

Figure 3.4 shows the range of values Įq in this case for then used to populate design equations. Such an approach has
roughness in the range of R1 (smooth, indentations less then been presented by Seidel and Collingwood (2001) where the
1mm) to R3 (grooves 4-10mm deep, >5mm wide, spacing 50- observations made by previous researchers have found a more
200mm) as defined by Rowe and Armitage (1987). The plot rigorous home. The method combines the effects of the
demonstrates the non-linear relationship between shaft friction following:
and rock strength. Further data is provided by Rowe and 1. Rock mass stiffness;
Armitage (1987) who suggested average values of shaft friction 2. Rock socket construction disturbance and cleanliness;
constants (for relatively small diameter sockets, typically 3. Socket geometry (length, diameter and roughness) (see
less than 0.6m diameter) as follows: Figure 3.5 below);
• c is equal to 0.4 for R1 to R3 sockets and 0.55 for R4 4. Rock strength (unconfined compression strength and the
sockets (R1 to R3 as above, R4 socket quality is defined as critical state angle of shearing resistance, φ'cv, of a fissure
grooves with depth and width greater than 10mm and as would be measured in a shear test)
spacing between 50 and 200mm).
• is equal to 0.57 for R1 to R3 sockets and 0.61 for R4
sockets

Figure 3.5 Illustrative model of pile rock socket displacement under


axial compressive load (from Seidel and Collingwood, 2001, after
Johnston and Lam, 1989)

These attributes will be considered in turn.


1. Beyond controlling the overall settlement in the rock mass,
rock mass stiffness has a significant effect on ultimate skin
friction by providing restraint to dilation on the pile/rock
interface. Dilation occurs as a function of the socket
roughness where shear displacement on the interface results
Figure 3.4 Relationship between shaft resistance and rock UCS for in radial strains in the surrounding rock mass and
socket roughness in R1 to R3 range (after Rowe and Armitage, 1984). corresponding increases in normal stress acting on the
Note use of qu for UCS of rock, ıci used elsewhere.
pile/rock interface. The constant normal stiffness value (K),
is related to the rock mass modulus (Em) and it is inversely
It is considered that such methods are useful for preliminary
proportional to socket diameter as shown in Equation 3.15:
design but must be used with caution where no comparable
experience is available (same rock, same pile size, same
construction process). The values shown above are best fit K = 2Em / [(1 + Ȟғғ).D] (3.15)
values and have been derived from load test data. Use of such
design constants within design codes such as Eurocode 7 Part 1 K constant normal stiffness parameter
(EN1997-1) could be seen to fail the requirement that the design Em rock mass modulus
process returns a resistance that is “accurate or erring on the Ȟғғ Poisson’s ratio
side of caution” (this requirement is not the same as a best D rock socket diameter.
estimate). The requirement results in the need to assess how a
“characteristic” skin friction should be derived. Assessment of It is the parameter K which controls the ability of the rock to
such characteristic values of skin friction may be obtained by resist dilation. Clearly, as seen in Equation 3.15, the larger the
adopting cautious estimates of both ci and c. However, even pile diameter the smaller the value of K. This is a similar effect
when this is done several of the Eurocode 7 National Annexes as seen for driven piles in sand above where the beneficial
(eg the UK national annex for Eurocode 7 Part 1: BSI, 2007b) effects of dilation are greatest for small diameter piles.
require that a further model factor be applied to the values to The rock mass modulus (Em) can be measured directly (eg by
acknowledge the uncertainty in the calculation and to arrive at dilatometer testing), or is more usually correlated with the rock
pile resistances in keeping with previous national practice and UCS, Rock Mass Rating (RMR) or rock mass quality (US
expectation. Army, 1994)
As noted above, the use of such empirical relationships is 2. The construction process used to form the pile and the
only acceptable when the design constants are available from cleanliness of the pile shaft are important in assessing the
comparable pile load tests, i.e. load tests on similar sized piles efficiency by which the pile shaft can transfer load to the
(socket size diameter and length) formed using similar ground. High quality construction leading to a clean shaft
equipment (drill bit detail and flush methodology) carried out in provides highest resistance while low quality construction
the same bearing stratum. Use of extrapolated data is inherently leaving smear results in low resistance. This variation in
dangerous. Where, as is likely to be the case with most large construction quality can be represented by the factor Șc as
diameter piles, good quality comparable pile load data does not discussed below.
exist then it is necessary to look to design methodologies which 3. The rock socket geometry is critical in assessment of shaft
are theory based, and which have been checked elsewhere capacity with rough sockets resulting in increased
against pile load test data as required by Method 2 in section 3.1 interaction between the pile and the surrounding rock (see
above. Appropriate design methods should allow for variability earlier comments on socket resistance above for R1-R3 and
in ground conditions, construction techniques and pile size by R4 roughness from Rowe and Armitage, 1987). The socket
means of readily measurable parameters. These parameters are
roughness is expressed in terms of change in radius of the
2892 B. Simpson et al. / State of the art Report: Analysis and Design

socket wall ( r). Seidel and Haberfield (1995a) present a bentonite or polymer usage is in keeping with manufacturer’s
theoretical design development of rock sockets specification and that the product is appropriate for ground conditions
incorporating a Joint Roughness Coefficient (JRC) and and usage for forming rock sockets.
fractal mathematics which better describe the effects of Construction method c

scale (both of asperity height and geometry) of rock socket Construction without drilling fluid:
roughness. Such discussion is outside the scope of this • Best practice construction and high level of 1.0
paper but is worthy of reference for future designers. construction control (e.g. socket side wall free of
Further data is presented in Seidel and Collingwood (2001) smear and remoulded rock)
where examples of roughness with rock UCS is presented as • Poor construction practice or low quality construction 0.3-0.6
control (e.g. smear or remoulded rock present on side
a first pass estimate. Whilst less common in practice, direct walls)
measurement of socket roughness with a profiling tool is Construction under bentonite slurry *1
also possible. • Best practice construction and high level of 0.7-0.9
4. The final consideration is associated with rock strength as construction control
represented by the intact strength (UCS) and fissure strength • Poor construction practice or low quality construction 0.3-0.6
(friction). Constant volume fissure strength (ϕ'cv) is used to control
assess the pile capacity when combined with socket Construction under polymer slurry *1
roughness and the normal effective stress acting on the • Best practice construction and high level of 0.9-1.0
construction control
pile/rock interface as a function of in-situ stress (small
• Poor construction practice or low quality construction 0.8
effect) and stresses mobilised due to attempted dilation at control
the socket to rock interface. For high strength rocks the
limiting strength may be that of the concrete rather than the
calculated (frictional) socket shear strength.
Seidel and Collingwood (2001) took the above
considerations and combined them in a dimensionless Shaft
Resistance Coefficient (SRC) that can be used to predict shear
stress mobilised at various displacements. The SRC is defined
as:

n Δr (3.16)
SRC = η c .
1 + ν 2rs

Șc construction method reduction factor as in Table 3.1.


n modular ratio (Rock Mass Modulus / rock UCS) = Em / ıci
ǻr mean roughness height of the rock socket, measured
directly or assessed from the asperity length and mean
asperity angle
rs rock socket radius
Figure 3.6 SRC and adhesion factor (Įq) (after Seidel and
The value of SRC can then be used to assess the appropriate Collingwood, 2001)
value of Įq thus allowing for rock socket adhesion to be
assessed from Equation 3.12 with allowance for cleanliness and Where rock socket embedment is less than 1 diameter, or
construction, pile diameter, rock stiffness and rock socket where the rock joints are not tight and clean, reduction in
roughness. Figure 3.6 shows the computed values for Įq with capacity is required. Where rock is of lesser quality it is
SRC. Further extensions of the work allow for assessment of necessary to accommodate the rock quality in assessment of
mobilised resistance versus displacement to build up a full base resistance. For this case attention to the nature of the rock
picture of pile loading. jointing and infill is required; the reader is referred to
The above comments have addressed limiting conditions in publications such as Rock Engineering (1994) by the US Army
the rock. It is also necessary to check the strength of the Corps of Engineers where a process based on rock mass
reinforced concrete pile element in axial, bending and shear considerations is provided. Where possible load testing should
modes as well as the limiting interface strength in the concrete be considered and to this end use of Osterberg cells provides a
adjacent to the rock socket. Such considerations require useful tool for testing components of base and shaft resistance
consultation with the appropriate structural design codes. independently.
Base resistance:
Base resistance may in rock sockets may be included in 3.3 Pile design – lateral loading
design when it can be demonstrated that cleanliness of the base
can be achieved and where the drilling and cleaning processes A literature review for papers and publications on laterally
have not altered the quality of the rock immediately below the loaded piles does not reveal the same levels of development as
socket base. When this is the case then rock socket base seen above for axial loading of piles. Recent distillations of
resistance can be expressed as a function of the unconfined design guidance for lateral loading include:
compressive strength of the rock, ıci, with an equivalent bearing • GEO (2006): assessment of single piles in soil with recourse
capacity factor as a function of rock quality and jointing. For to Broms (1964) and Brinch Hansen (1961);
massive or tightly jointed rock (clean joints) with a socket depth • Turner (2006): assessment of single piles in rock with
at least one diameter into equivalent quality rock, Rowe and highway structure loading, recourse to Reese (1997);
Armitage, (1987) suggest limiting ultimate base resistance, qb, • API (2000); and
to: • Randolph et al (2005).
qb = 2.5 ci (3.17) Such documents provide useful summaries of current design
practice but do not reveal significant development in design
Table 3.1 Construction method reduction factors (after Seidel and processes. Turner (2006) and API (2000) both present analysis
Collingwood, 2001). Note *1: Care is required to ensure that the methods utilising “p-y” curves to model the soil support to
B. Simpson et al. / State of the art Report: Analysis and Design 2893

laterally loaded piles. These p-y (force – displacement) springs 3.4.1 Foundation reuse
have limiting values (pur) which can be calculated according to Reuse of foundations is a critical consideration for urban
classical solutions such as that proposed by Brinch Hansen regeneration and is well documented in Butcher et al (2006a)
(1961). For limiting values of lateral resistance, Martin and and papers within the 2006 conference “Reuse of Foundations
Randolph (2006) report more recent work for deeply embedded for Urban Sites” edited by Butcher et al (2006b). Foundation
piles in clay soils (undrained behaviour only) while Reece reuse has the potential to:
(1997) provides a calculation for ultimate lateral resistance for • reduce cost and construction programmes;
piles in rock as follows: • reduce the impact on archaeological resources;
• reduce disturbance to contaminated ground; and
pur = r ci
D (1 + 1.4 xr / D) • 5.2 r ci
D (3.18) • avoid disturbance to existing underground services and
structures.
where
These benefits are in part countered by the need for:
is a reduction factor and is equal to 0.33 for an RQD of
r
100% increasing linearly to 1.0 for an RQD of 0%. Values • careful and advanced planning;
of r above 0.5 should be used with caution and it is • additional testing and investigation;
recommended that a limit on r of 0.5 is used where • more complicated design; and
comparable load test data is unavailable. • potential increased costs in design and construction of pile
xr is depth below top of rock caps and transfer structures.
Published data for p-y curves are generally for use with The main tenets for foundation reuse are presented below
single piles; they do not accommodate behaviour of piles in and may be traced from Butcher et al (2006).
groups where there may be significant interaction between piles Reuse of existing foundations:
as well as structural restraint at the pile head influenced by pile • Foundation reuse is not a decision that a designer can take
cap stiffness. Hence, for pile groups recourse to p-y curves alone. The decision to reuse, while laudable, must be
assessed from single pile load test data is not acceptable for agreed upon by the client, the client’s insurers and the
detailed design and alternatives are required. Alternative building control authorities (City engineer or similar). In
approaches include the following: specific cases, other interested parties may be architects and
• Preliminary design can be carried out by using p-y curves to structural engineers (to ensure correct coordination of
generate single pile behaviour. The results from the single superstructure to substructure and to allow space for transfer
pile model can then be used to calibrate pile group computer structures), future tenants of the structure and their insurers,
codes such as PIGLET (User manual: Randolph 2003b) the list is long. The geotechnical designer can lead this
with single or very widely spaced piles. Such calibration is process by documenting the cost, programme and
particularly important with layered soil where the relatively environmental risks and benefits of foundation reuse
simple input to pile group programs (linearly increasing thereby achieving sign-on from those parties who will be
stiffness with depth) does not lend itself to the stepwise most influential in a successful outcome. The process
varying stiffness values of layered soils. The resulting adopted by the designer which results in pile reuse should
assessment of spring data can then be used for assessment be one where reasonable skill and care is demonstrated and
of pile group behaviour. Upper and lower bounds of input this should be part of the designer’s contract. It would be
parameters should be obtained for the single pile to allow impossible for a designer to provide a fit for purpose
pile group behaviour to be investigated. guarantee for pre-existing structures regardless of the
• For detailed design, especially where there is no case diligence of his investigation and design. As such ultimate
history data or where pile cap interaction with the ground is risk, accepting reasonable skill and care by the designer will
important, use of 3D finite element (or similar) computer reside with the client and, hopefully, his insurers in the final
programs will provide the optimum analysis. It is however place.
noted that offshore industry practice (e.g. API, 2000) • Foundation reuse should not detrimentally impact on the
usually reverts to methods using p-y and t-z/q-z (vertical performance of the structure being placed on it, nor should
stiffness) for pile group analysis, such methods are included it limit the function of the new structure. If there is
in commercial computer programs with varying degrees of reasonable doubt relating to the performance of a
complexity (structural stiffness of pile cap, non-linear foundation when reused then these doubts should be
springs etc). These codes do not provide full soil structure addressed by means of measurement or redesign. All risks
interaction and as such require a degree of experience in and contingency plans must be agreed by risk holders before
their use for large groups of closely spaced piles is required. implementation of a construction reliant on foundation
reuse.
3.4 Developments in pile design – environmental • When reloaded, foundations will be stiffer than on initial
considerations. loading and probably stiffer than newly built foundations.
However, when foundations are loaded beyond historic
What constitutes good design? Historically a good design was maximum loads there will be a marked non-linearity in the
one that provided the required function at minimum cost, was load-settlement curve when compared to newly built piles
constructed within the programme requirements and resulted in which, while less stiff, will have a more uniform stiffness
minimum disruption to third parties, these issues are addressed up to working load (less non-linearity). In the case of a
in Section 2 in terms of code requirements. The question is foundation solution using both new and existing foundations
“can we do better than this?” Recently published books and the difference in stiffnesses during loading should be
papers on foundation design for buildings show two significant investigated to prevent structural over-load of piles or other
developments that should be considered in providing a “total” elements of the sub-structure as a result of load
design: redistribution.
• reuse of existing foundations; • At the limit, the ultimate load capacity of old piles,
• planning for new foundations to be reused; and especially those which have been loaded for significant
• incorporation of piled foundations into the building’s periods, is seen to be larger than similarly dimensioned new
heating and cooling systems. piles. This potential benefit must be balanced against non-
linear settlement impacts as described above.
2894 B. Simpson et al. / State of the art Report: Analysis and Design

• Structural performance of the reused foundations should be a means of improving the green credentials of new build
fully understood. As-built records of these foundations are structures.
valuable but must be confirmed in terms of geometry (pile
diameter, length, reinforcement quantities and depths) as
well as material strengths and durability. Where there are
no as-built records, an enhanced level of testing will be
needed. It may be that insufficient confidence can be
obtained related to the existing foundations to allow reuse.
The existing structure above foundations planned for reuse
should be checked for distress (did the foundations perform
as would be anticipated?); load intensity (what load has
been applied to the piles?); and geometry (does the pile
layout match the substructure / superstructure?). If these
investigations suggest a building that is at odds with the
anticipated foundation then further investigation will be
required.
• When foundation reuse is to be adopted the existing
structure must be sympathetic to foundation reuse. Access
to piles must be possible to check location, geometry and Figure 3.7 Pile with tubing installed prior to pile construction. The
materials. Can load testing be carried out using the mass of tubing is often high density polythene plastic pipes of 20 or 25mm
the existing building as reaction? Does demolition cause diameter. The circulating fluid typical comprises either water, water
damage to the structure of existing piles (ground movement and an anti-freeze solution or a saline solution depending on the
operational temperature range
causing cracking of piles in bending or tension)?
The above list provides the basic considerations for reuse of
Brandl (2006) summarises the current considerations for
existing foundations. The following list presents the basic
energy pile design in a detailed presentation of his 2001
requirements that present day designers and contractors should
Rankine Lecture, illustrating good practice by means of case
satisfy in order that reuse of new foundations is a viable option
history data and analysis. Energy piles have the ability be used
in the future. Planning for reuse:
cyclically by extracting heat from the ground when the building
• Provision of documentation for the future is valuable and a is cooler than required and providing “coolth” when the
relatively easy task. A “close-out report” should include: building is warmer than required. The “energy” part of energy
full details of ground conditions, foundation design piles is usually subservient to their load carrying capacity, so
calculations, design and as-built layout and geometry care to prevent the cyclic cooling and heating of the piles and
(dimensions and depth) and materials; full records of testing surrounding ground from causing damage to the soil matrix and
(both pile load testing and material testing) as well as pile pile structure is a prerequisite. Brandl describes the process
installation logs and non-conformance reports (with any where the heating-cooling cycle is annual and this is the most
mitigations recorded). The designer’s/contractor’s typical requirement for domestic situations. It is however noted
programme and costs should allow both the time and that the heating-cooling cycle can be daily allowing
resources to complete such a report; without such financial significantly higher temperature differentials to be achieved.
consideration a “close-out” report will not be provided in a Such systems require more sophisticated management and are
commercial world! In time the report will have a value way more appropriate to commercial buildings where peak daily
in excess of the cost of production. The report should be conditions require cooling during day time and heating during
deposited with the building records maintained by the night time (e.g. office buildings with significant computer usage
owner. Other documents to be preserved include the factual or retail with high energy lighting). Where energy piles are
ground investigation and design reports as well as the piling operated with conductive fluids at less than 0°C, greater care in
specification and contractor’s method statements. the design will be needed to ensure that temperatures less than
• Where possible, monitoring of building performance is a 0°C are restricted to the pile structure and do not reach the
useful demonstration of foundation performance; this is ground. In such situations accurate modelling of thermal
usually only viable for larger structures. Such monitoring properties of materials and the construction geometry of a the
includes settlement monitoring in its simplest form but may piles are important; Figure 3.8 shows a model simulation of an
also include load monitoring in piles (strain measurements energy pile after 15 hours operation with a circulating fluid at -
usually) or pressure monitoring below rafts. The 5°C. It is clear that the pile edge temperature has dropped to
monitoring data will validate the design and construction to 1°C. Further operation at a temperature of -5°C would result in
working conditions and will provide subsequent owners local, increasing to general, freezing of the ground around the
with data to assess future foundation performance. pile and then potentially catastrophic thawing. The rate and
duration of energy extraction from the ground should be
3.4.2 Energy piles analysed using thermodynamics to check that the ground does
Energy piles are conventional load bearing piles which have not freeze (and detrimentally thaw at a later date) and that the
been equipped with ducting to carry a circulating thermal fluid, pile structure is not damaged due to temperature induced
usually attached to the reinforcement cage as illustrated in stresses. An investigation of the latter of these issues has been
Figure 3.7. The ducting allows the piles to be used as heat carried out by Bourne-Webb et al (2009) where thermal changes
exchange elements as well as being load carrying structures. where induced on an energy pile during (maintained load)
The thermal fluid in the ducting is usually allied to heat pumps preliminary load testing. The pile load test demonstrated a
to amplify temperature differences. This allows the energy piles relatively simple model associated with skin friction on the pile
to be used as part of the building management system (BMS) by shaft being mobilised to resist both applied pile load and
providing all or part of the heating / cooling demand of the thermally induced strains. The result is modified stresses in the
building. pile varying with depth and the need to account for these
The use of energy piles has been widespread in continental stresses in structural design of the pile. Bourne-Webb et al also
Europe since the 1980’s. Elsewhere, it is gaining importance as
note that the peak skin friction mobilised during the thermal test
B. Simpson et al. / State of the art Report: Analysis and Design 2895

was marginally lower than the expected design value. Whilst 4 EMBANKMENTS AND SLOPES
the difference between mobilised and expected skin friction was
not particularly large (17%) it is further evidence that the design
of energy piles must consider how thermal effects will impact 4.1 Introduction
on static design, especially where partial factors on skin friction
are low. More maintained load testing is required with thermal In this section of the paper, recent developments in the analysis
variations superimposed on the pile. and design of embankment sand slopes is described, with a
particular emphasis on Japane≤se practice. Slope stability is a
very important issue in a seismically active region, resulting in
active development of both empirical and analytical methods of
design.

4.2 Basic elements of design – embankments

4.2.1 Classification of embankments and items to be


considered in the design of cut slope
Figure 4.1 shows the main types of embankments considered in
this paper: road embankments, railway embankments, reclaimed
areas for housing lots and buildings, river dikes and sea dikes.
Design methods for dams are more severe and are not included
here. Items to be considered in the design are: fill material,
density of embankment, gradient and height of slope, and
Figure 3.8 Section of a pile model with circulating fluid at -5°C after drainage facilities. In addition, seepage of water must be
15 hours operation. considered in river and sea dikes.

In general, thermal piles are usually associated with


conventionally bored cast-in-place piles, although driven
reinforced concrete piles with integrated absorber pipes are
occasionally used. CFA piling, requiring the plunging of the
thermal ducting and reinforcement cage into wet concrete, is not
generally advised due to the possibility of damage to the
thermal pipes during cage installation into wet concrete.
Nevertheless it is anticipated that development of thermal CFA
piles will continue due to the economic benefits of such piles.
Construction modifications, beyond that which is needed for
structural considerations, include stiffening the reinforcement
cage to help support and protect the thermal ducting for systems
with high energy demand and the development of plunging
mandrels to install the ducting below the cage in CFA piles Figure 4.1: Types of Embankments
where lower energy demands are appropriate. For details on
design the reader is referred to Brandl (2006) in the first place 4.2.2 Fill material
and thereafter to 2009 issues of Géotechnique for the Selection of appropriate fill material is important because bad
Symposium-in-Print on “Thermal behaviour of the ground” (in material may cause poor workability, settlement or instability.
preparation at time of this article being submitted). Good fill materials have the following properties:
1. Placing and spreading, and compaction of the material is
3.5 Conclusions – Piling developments easy.
2. Shear strength and bearing capacity are high, and
compressibility is low.
It is hoped that the above sections show how the field of pile
3. Water intake swelling is low.
design, analysis and construction continue to develop.
4. They are stable against erosion and shear strength does not
Examples of offshore driven piles in sand and rock socket piles
decrease due to saturation.
have been presented with shorter comments on axial design of Well graded gravelly or sandy soils satisfy these conditions
piles by means of the Menard Pressuremeter. The common and are recommended for use as fill material. On the contrary,
thread that these two examples exhibit, especially for shaft expansive soils such as bentonitic or solfataric soil (sulphurous
friction, is that of detailed assessment of data within a volcanic soil) and highly organic soil cannot be used to
framework that is related to the model by which skin friction is construct embankments. Attention is necessary to the following
mobilised. Care must be taken with extrapolating these special soils:
methods to pile sizes beyond their respective data bases even a. Volcanic cohesive soil: clay originating from volcanic ash.
though both methods include pile diameter effects. Void ratio and natural water content are very high and unit
The second topic addressed in this chapter is that of weight is very low. Therefore, trafficability is low, and the
sustainable design. The need to achieve not only economic but embankment may suffer from slope failure or long-term
sustainable designs is ever pressing. Such pressures should settlement.
facilitate cross fertilisation of skills and knowledge from other b. Some sedimentary soft rocks such as mudstone, shale and
areas of engineering as well as encouraging designers and tuff: these may suffer slaking and water intake swelling
contractors to look to the future in all that they do. The move to after filling.
include more than one function in the design of a pile is
laudable as are the benefits of reusing existing piles and 4.2.3 Density of embankments
providing documentation to encourage future reuse of current Appropriate placing and spreading, and adequate compaction
piling projects. must be conducted during construction. In general, the
2896 B. Simpson et al. / State of the art Report: Analysis and Design

compaction control standard is decided before construction. in which u is the pore water pressure including excess pore
Several compaction control methods are available as follows: water pressure due to liquefaction, and kh: is the horizontal
a. Compaction control by density based on degree of seismic coefficient.
compaction. Fellenius’ method is used for simplicity, although its
b. Compaction control by degree of saturation. shortcomings are acknowledged (Lambe & Whitman 1979).
c. Compaction control by strength of deformation. One simple method to estimate excess pore-water pressure due
d. Compaction control by compaction method. to liquefaction (Japan Road Association, 1986) is to assume that
Ru is 1.0 when FL≤1.0, where, Ru is the ratio of excess pore-
4.2.4 Gradient and height of slope water to effective overburden pressure. When Ru = 1.0, the
In general, failure of an embankment occurs for one of two entire weight of overburden is taken on the water pressure.
main reasons: a) failure due to insufficient strength of the A new design concept to introduce performance-based
foundation ground during construction of the embankment, and design (PBD) has been developed recently. In the PBD,
b) failure triggered by heavy rains or earthquakes after deformation of an embankment must be evaluated. Several
construction. In the case of dams, their height and inclination of methods studied recently are:
slope are evaluated by conducting slope stability analyses. 1. Dynamic response analysis (FEM)
However, in case of road and railway embankments, 2. Static residual deformation analysis (FEM)
inclinations of their slopes are usually designed empirically 3. Newmark’s method (Newmark 1959)
without conducting slope stability analyses, because their length
are very long. In the empirical approach, the appropriate 4.2.5 Drainage facilities
inclination of a slope is decided by the height and the material Drainage is very important to construct safely and maintain
of the embankment. The empirical design method is not uniform stable embankments. Usually seepage water and surface water
throughout the world because fill and ground conditions are must be cut off from surrounding area into the construction site
quite different in each country. For example, road embankments temporarily, during the construction of embankments. In the
are designed based on Table 4.1 in Japan. design of permanent drainage, two forms of drainage must be
considered as shown in Figure 4.2: (a) drains for surface water,
Table 4.1 Empirical design method for road embankment in Japan including subsurface drainage and slope surface drainage, and
(Japan Road Association, 1999) (b) facilities against seepage water into the embankments, such
Fill Material Height of Inclination as horizontal drains, horizontal blankets, drainage pipes or
embankment of slope gabions.
Well graded sand, Less than 5m 1:1.5 to 1:1.8
Gravel, Gravel with 5m to 15 m 1.1.8 to 1:2.0
fines
Poor graded sand Less than 10m 1.1.8 to 1:2.0
Less than 10m 1.1.5 to 1:1.8
Rock, Muck
10m to 20m 1.1.5 to 1:1.8
Sandy soil, Hard Less than 5m 1.1.5 to 1:1.8
clayey soil, Hard clay 5m to 10m 1.1.8 to 1:2.0
Volcanic clay Less than 5m 1.1.8 to 1:2.0

If the embankment is high and/or the material is soft, low


inclination is selected. However, slope stability analysis is
necessary if conditions for the foundation ground and/or the
embankment are bad as follows:
a. the underlying ground is soft. Figure 4.2: Drain facilities for embankments
b. the embankment is constructed in landslide area.
c) the height of the embankment is greater than the heights Special care for drainage is necessary to the following
listed in Table 4.1. locations:
d. the fill material is bad, such as high moisture clay or a. Small rivers or boundaries between cuts and embankments
volcanic ash. where flows of surface water concentrate
e. houses exist near the embankment that could be damaged if b. Valleys, and boundaries between cuts and embankments
the embankment is deformed. where much spring water emerges.
For stability, circular slip surface analysis is widely used.
However, compound slip surface analysis must be conducted if 4.3 Basic elements of design- Cut slopes
a thin soft layer exists beneath the embankment or if the surface
of the foundation ground is inclined. Procedures to estimate
safety factor by finite element methods have also been 4.3.1 Items to be considered in the design of cut slopes
developed recently. In the design of cut slopes, slope gradient, drainage facilities,
In seismic design, the seismic force is added in the stability slope protection works and maintenance works must be
analysis. Moreover, the effect of excess pore water pressure is evaluated.
considered if the foundation soil and/or fill material are
liquefiable. One formula considering the excess pore-water and 4.3.2 Design of slope gradient
derived based of Fellenius’s method is as follows: If the natural ground of the slope is composed by collapsible
soils such as shown in Table 4.2, soil investigation, soil tests
(4.1) and slope stability analyses are necessary to design appropriate
Fs =
¦ [cl + {(W − ub) cos α − k W sin α }tan φ ]
h
slope gradients. If there is a flow of water in the ground, slope
§ h · stability analysis is necessary even in normal soils.On the
¦ ¨©W sin α + r k W ¸¹
h
contrary, if the slope is protected from water flow, the gradient
of the slopes composed by normal soils is often designed by
empirical methods without conducting soil investigation, soil
tests and analyses. In the empirical methods several factors,
B. Simpson et al. / State of the art Report: Analysis and Design 2897

such as hardness of soils, height of cut slope, are considered. The planting treatment is conducted by spraying of seeds,
The empirical design method is not uniform throughout the pasting of special mats which are processed by soils and seeds,
world because soil conditions of slopes are quite different in or turf work. In the slope protection by concrete pitching etc,
each country. For example, road cut slopes are designed based four types of methods are in use:
on Table 4.3 in Japan. a. protection against weathering and erosion: guniting
(spaying mortar), shotcreting (spraying concrete), stone
Table 4.2 Collapsible soils facing,
Soil type Typical soils
b. protection against flow of soils due to erosion and spring
water: gabion
Weak for erosion Welded tuff, weathered granite
c. protection against sliding: concrete pitching, concrete crib
Unconsolidated or weathered Talus, Volcanic soil, Colluvial
work.
deposit
Weathering speed is high Mudstone, Tuff, Shale, Slate,
Serpentine
Fissured Shale, Serpentine, Granite,
Andesite
Fissures forms dip slope Slate
Sandwiching soft layers Fault clay, Post landslide or
collapsed site

Table 4.3 Empirical design method for road cut slopes in Japan (Japan
Road Association, 1999)
Soil of cut slope Height of cut Inclination of
slope cut slope
Hard 1:0.3 to 1:0.8
rock
Soft rock 1:0.5 to 1:1.2
Sand Not dense and More than 1:1.5 Figure 4.3: Drainage facilities for cut slope
poor graded
Less than 5m 1:0.8 to 1:1.0 4.3.5 Retaining walls
Dense
Sandy 5m to 10m 1:1.0 to 1:1.2 If the cut slope is steeper than the stable inclination,
soil Less than 5m 1:1.0 to 1:1.2 construction of a retaining wall at the toe of the slope is
Not dense
5m to 10m 1:1.2 to 1:1.5 appropriate. If the wall is high, such as more than 5m, stability
Dense or well Less than 10m 1:0.8 to 1:1.0 of the wall against earth pressure must be checked by
Gravelly graded 10m to 15m 1:1.0 to 1:1.2 calculation; for lower walls, empirical design based on
sand Not dense or Less than 10m 1:1.0 to 1:1.2 experience may be used.
well graded 10m to 15m 1:1.2 to 1:1.5
Clay Less than 10m 1:0.8 to 1:1.2 4.3.6 Maintenance
Gravelly Less than 5m 1:1.0 to 1:1.2 Ground and protection work of cut slopes becomes deteriorate
clay 5m to 10m 1:1.2 to 1:1.5 with age owing to several factors. The most important factor is
rainfalls and seepage from bedrock, for which appropriate
For slope stability, compound slip surface analyses or treatment is necessary based on periodical site inspection of
circular slip surface analyses are commonly used for shallow seepage water and drainage conditions.
slides and deep slides, respectively. In seismic design, seismic
force is added in the stability analysis. However, this is seldom 4.4 Basic elements of design - natural slopes
to be conducted. In the new design concept to introduce
performance-based design, deformation of cut slope may be
evaluated by dynamic response analysis or Newmark’s method. 4.4.1 Classification of landslides
But more studies are necessary to introduce these methods. There are many types of landslides of natural slopes. If the
landslides are classified by speed of sides, landslides are
4.3.3 Drainage facilities classified into two groups:
For cut slopes, several kinds of facilities must be equipped for 1. Rapid landslides
drainage as schematically shown in Figure.4.3. These drains are Steep slopes slide rapidly during heavy rainfalls or
classified into two groups: (a) facilities against surface water, earthquakes. Slipped masses fall down to the foot of slopes.
such as catch ditches at the tops of slopes, drain ditches on Surface slides of weathered soils are dominant, but if some
berms and drain channels along slopes, and (b) facilities against deep layers are week, deep slides may occur during
seepage water and spring water, such as drain conduits at the earthquakes. For example, the maximum depth of the slip
toes of slopes, horizontal drainage pipes and drainage wells. surface of Ontake Slide which occurred during the 1984
Naganoken-seibu earthquake was about 180m. Stability of
4.3.4 Slope protection works against sliding, weathering or natural slopes is generally judged roughly by several factors
erosion such as angle of slope, existence of unstable rocks, histories
Slope protection works are classified into two groups: planting of slides. Stability analyses are conducted in special cases
treatment and slope protection by concrete pitching, reinforcing only. In the analyses, stability of the slides can be evaluated
grids, sprayed concrete etc. The latter is applied only to the based on peak strength of the soils. However, prediction of
following conditions because of high cost and bad appearance. unstable slopes over wide areas during future rains and
a. The soil of the slope is too hard or acid for vegetation. earthquakes is not easy.
b. The slope is unstable against sliding. 2. Slow landslide (land creep)
c. Protection against erosion is necessary because of running Gentle slopes may creep slowly and periodically due to
surface water from rainfall or springs causing slope failure. seasonal change of water level. Sliding may occur
particularly in spring if the slope is affected by melting
2898 B. Simpson et al. / State of the art Report: Analysis and Design

snow, possibly sliding a few metres each year on weak


layers, such as weathered pumice and sandstone. Stability of
the slide can be evaluated by slope stability analyses based
on residual strength. As the slides occur periodically, slide
prone sites can be identified easily, and countermeasures
can be applied.

4.4.2 Countermeasures
Two types of countermeasures against landslides may be
Figure 4.4: Deformation of an embankment constructed by EPS method
introduced: control works and prevention works. In control
due to the weight of covered soil
works, the stability of a slope can be increased by controlling
ground water level or the shape of the slope. In prevention
4.5.3 Judgment of instability of embankments by field
works, failure of the slopes is prevented by constructing some
observation
countermeasures against sliding. Some countermeasures are
If the ground is very soft, stability of an embankment is
listed in Table 4.4.
evaluated not only before construction but also during the
Table 4.4 Countermeasures against landslides
construction of the embankment. Settlement and spreading of
the toe of the embankment are measured and may be plotted on
Type of Countermeasure methods a control diagram such as Figure 4.5.
countermeasure If the plotted point exceeds a failure (critical) line, it is
Control works Surface water drainage work Groundwater judged that the embankment is becoming unstable, and the
drainage work construction cannot be continued. An appropriate measure, such
• Ground water cut-off wall as waiting several days or constructing counterweight fill, is
• Soil removal work applied. Many control diagrams have been proposed based on
• Counterweight fill method
case studies of failed embankments.
Preventing works Retaining wall
Soil reinforcement
Ground anchorage
Pile works, Shaft work

4.5 Recent developments- embankments

4.5.1 Effective use of construction generated soils


Recently reconstruction of buildings and new construction of
underground structures have increased in urban areas.
Consequently, the availability has increased of soils from
excavations and broken brick and concrete from demolition of
buildings. These materials are difficult to dispose of because of
lack of landfill. So it becomes necessary to use these materials
for fill material or other purposes. Often, excavated soils are not
suitable for fill materials because their water content is too high
and/or they are unsuitable mixtures of coarse and fine grains.
Then special treatments are conducted as follows:
a. lowering of water content by special bags or sandwich
methods;
b. treatment of soil with air foam, fibers, cement or lime.

4.5.2 Lightweight embankments Figure 4.5: Diagram to judge instability of embankments (proposed by
A recently developed method to prevent large settlement due to Matsuo and Kawamura 1977)
construction of embankments on very soft ground is to use light
weight material. EPS (Expanded polyestyrene) construction 4.5.4 Introduction of performance-based design in seismic
block, lightweight treated soil with air foam (see, for example, design
Tsuchida et al 2001), lightweight treated soil with EPS beads, or The need to evaluate not only the safety against sliding but also
fly ash may be used as the light weight material. The use of EPS deformation of embankments was recognized after the 1995
method has become especially popular recently. In this method, Kobe earthquake. In river dikes, the aim is to protect from
EPS blocks are stacked and covered by soils. If the weight of flooding, so the critical requirement of the dikes is prevent
the soils for cover is heavy, settlement at the toe of embankment overflow of river water, as schematically shown in Figure
occurs as schematically shown in Figure 4.4. Therefore 4.4(a). In road embankments, emergency vehicles must run just
attention is necessary not only the weight of EPS but also that after earthquakes. For example differential settlement of
of the covered soils. Floating due to buoyancy and disturbance approaches to bridges from embankments must be within the
due to strong wind must be prevented during construction. appropriate value for the vehicles, as shown in Figure 4.4(b).
Moreover, EPS may be burned if there is a fire due a to traffic Thus, differential settlement must be one of the critical
accident, and it can also be damaged by petrol spillage. conditions in design of road embankments. One more reason
why the evaluation of deformation of embankments must be
introduced in seismic design, is that the calculated safety factor
against sliding, Fs, is apt to be lower than 1.0 under the Level 2
earthquake motion, even though the ground is medium dense.
Therefore Fs cannot be used in the design under Level 2
shaking. Here, Level 2 shaking motion is defined as the
maximum shaking motion. Level 1 shaking motion is defined as
B. Simpson et al. / State of the art Report: Analysis and Design 2899
Table 4.5: Damage level for railway embankment (Railway Technical
Institute 1999)
Deformat Damage level Settlement, S (roughly
ion level speaking)
1 No damage None
2 Slight damage S<20cm
Figure 4.6: Critical conditions for river and road embankments 3 Medium damage 20cm҇S <50cm
(restoration is available
the shaking motion which occurs two or three times during the with emergency repairs)
lifetime of a structure Recently, allowable settlements for super 4 Severe damage (long term S҈50cm
levees (Japanese River Association, 1997), railway restoration is necessary)
embankments (Railway Technical Institute, 1999) and river
levees (River Bureau, 1995) have been introduced in Japanese Table 4.6: Damage level for differential settlement between abutment
design manuals and guidelines. A “super levee” is an extended and embankment (Railway Technical Institute 1999)
area of raised ground behind the main levee, on which Deform- Damage level Differential settlement
construction may take place as shown in Figure 4.7. In the ation between abutment and
manual for super levees, allowable settlements are 50 cm for the level embankment, Sd (roughly
top of the levee and face of the back slope, and 20 cm for the speaking)
ground on the super levee. As the super levees are used for
1 No damage None
residential areas, similar safety as urban areas is necessary.
2 Slight damage Sd<10cm
Hence these allowable values were introduced in the design
3 Medium damage 10cm҇S d<20cm
manual.
(restoration is available
with emergency repairs)
4 Severe damage (long Sd҈20cm
term restoration is
necessary)

Table 4.7: Relationship between Fs and settlement (Public Works


Research Institute 1997)
Safety factor of slope, Fs Settlement (maximum)
Fsd(kh) Fsd(ǻu)
1.0<Fsd 0
0.8<Fsd<1.0 0.25H
Fsd0.8 0.6<Fsd<0.8 0.5H
Fsd<0.6 0.75H

Figure 4.7: Difference between normal and super levees

In the guideline for railways, damage levels of deformation


of embankments are classified into four grades as shown in
Table 4.5 and Table 4.6 for embankments and approaches to
bridges, respectively. If the settlement is more than 50 cm it can
be judged that long term remedial work is necessary. On the
contrary, it can be judged that the damage is slight if the Figure 4.8: Definition of allowable settlement for river dike (River
settlement is less than 20 cm. Therefore allowable settlements Bureau 2007).
under Level 1 and 2 earthquake motions were decided as 20 cm
and 50 cm, respectively as shown in Table 4.5. Recently,
seismic diagnosis of existing river dikes has been conducted in
Japan. In the diagnosis, allowable settlement is defined as
shown in Figure 4.8. It is recommended that the level of a river
crest after an earthquake must be higher than the level of the
mean monthly highest water plus wave height.
In the estimation of liquefaction-induced deformation of
structures, three types of methods are available: empirical
methods, static analyses and dynamic analyses. One empirical
method is introduced in the design manual for river dikes, as
shown in Table 4.7. In this method, settlement of a dike is
estimated by the safety factor of the slope, Fs. Two values of Fs
must be calculated: Fs (kh) which considers the seismic
coefficient, and Fs(ǻu) which considers excess pore water
pressure due to liquefaction. Then the settlement is estimated by
the lower FS. The relationship shown in Table 4.7 was derived
from the correlation between settlement of damaged dikes and Figure 4.9: Relationship between Fs and settlement ratio of river dike
Fs during several past earthquakes as shown in Figure 4.9. (Public Works Research Institute, 1997)

Figure 4.10 shows another empirical relationship between


settlement of dikes of Kiso, Nagara and Ibi Rivers during the
1944 Tohnankai earthquake, and liquefaction potential, PL at the
damaged sites. As show in this figure, the settlement increased
2900 B. Simpson et al. / State of the art Report: Analysis and Design

with the value of PL. PL is calculated by the following formula In dynamic and static analyses, several methods have been
together with Figure 4.11 (Iwasaki et al., 1978). proposed and applied to estimate the deformation of
embankments. A technical committee organized by the Japanese
20 (4.2) Institute of Construction examined the efficiency of these
PL = ³ (1 − FL )(10 − 0.5 z ) dz analytical approaches. In the examination, two dynamic
0
methods of analysis, LIQCA and FLIP, and two static methods,
where FL is the safety factor against liquefaction, equal to the ALID and Towhata’s method, were applied to seven actual river
ratio of undrained cyclic shear strength to cyclic shear stress. dikes which were damaged and non-damaged during the 1993
The term (1-FL) is set to 0.0 if FL>1.0 Hokkaidonansei-oki earthquake and 1995 Hyogoken-nambu
earthquake. LIQCA and FLIP are two-dimensional effective
stress computer codes developed by Oka et al. (1999) and Iai et
al. (1992), respectively. ALID is a simplified method using
static FEM developed by Yasuda et al. (2003) by assuming that
residual deformation would occur in liquefied ground due to the
reduction of shear modulus. Towhata’s method was developed
based on a minimum energy principle. Figure 4.13 shows the
comparison between the calculate dike settlements and the
observed settlements. Settlements estimated by the empirical
approach shown in Table 4.7 are also compared in the figure.
The predicted settlements by the analytical approaches agree
fairly well. Figure 4.14 shows the analysed deformation by
ALID at a severely settled dike. These methods were also
applied to models with countermeasures tested on a shaking
table apparatus, to demonstrate the applicability of the
analytical methods to the dikes with countermeasures.

Figure 4.10: Relationship between PL and settlement of river dikes


(Nakamura and Murakami 1980)

Figure 4.13: Comparison between calculated and observed settlements


(Sasaki et al 2004)
Figure 4.11: Definition of PL (Iwasaki et al 1978)

For railway embankments, a relationship among settlement,


height of embankments, density, number of cycles and
liquefaction potential, PL is prepared to estimate the settlement
of embankments as shown in Figure 4.12. This relationship was
derived from shaking table tests.

Figure 4.14: Analysed deformation by ALID (Shiribeshi-toshibetsu


River, No.1) [8]. Height of river dike: 5.3m. Settlement: 2.3 m.
(Yasuda et al 2003)

4.5.5 Development of strengthening techniques of existing


embankments (partially quoted from Yasuda (2007)
If the surface soil of the foundation ground is very soft clay or
peat, soil improvement is necessary before construction of the
new embankment. Many improvement techniques have been
developed such as sand drains, paper drains, preloading, and
cement-mixing methods. Recently, it has become necessary to
develop soil improvement techniques for existing
Figure 4.12: Relationship between PL and settlement ratio for railway embankments. In particular, it is desired to develop appropriate
embankments (Sawada et al 1999) countermeasures for the embankments on sandy ground which
is liquefiable due to earthquakes.
B. Simpson et al. / State of the art Report: Analysis and Design 2901

If the soil of an embankment liquefies, it is necessary to Table 4.8: Factors introduced in visual inspection to evaluate seismic
prevent the flow of the liquefied soil. Lowering the water level stability of existing fill slope
in the embankment by installing horizontal pipe drains may Item Factors introduced in visual inspection
prevent liquefaction. In contrast, if the soil under an Fill Height of fill, inclination of slope, Banking
embankment liquefies, large settlement of the embankment material, Type of slope protection work,
occurs due to horizontal movement of the ground below. In this Deformation of slope, Springwater, Drainage
case, restricting the movement by some technique, such as the well
installation of underground walls at the toes of the embankment, Retaining wall, Type of retaining wall, Type of weeping,
can reduce the settlement. Several techniques which have been house Deformation of retaining wall, Springwater,
applied to existing embankments recently are schematically Deformation of foundation of house
shown in Figure 4.15. The Tokaido Shinkansen Railway Topography Catchment, Land use on the slope
(Japanese Bullet Train) runs over areas where liquefaction is
possible during future earthquakes. As liquefaction may damage Several evaluation methods based on these factors have been
railway embankments, a sheet-pile enclosure method has been developed in Japan, which are introduced in the guideline. For
developed to protect them as illustrated in Figure 4.15(a) example, a rough check sheet by visual inspection for an
(Japanese Geotechnical Society 1998). The 1995 Kobe existing retaining wall prepared by Yokohama City is shown in
earthquake caused extensive damage to the Yodogawa dike. Table 4.9.
The deep mixing method was applied to the embankment of the Step 3: Soil investigations and laboratory tests
Arakawa River embankment in Tokyo, as shown in Figure For the selected unstable slopes, detailed soil investigation
4.15(b). A loose sand layer where liquefaction was anticipated and laboratory tests are necessary to evaluate the seismic
was 3 to 6 m thick. It was planned to use the deep mixing stability of the slopes more precisely.
method for a width of 10m and a depth of 24m to stabilise the Step 4: Analyses on slope stability or deformation
foundation ground against external seismic forces, which are Based on soil investigations and laboratory tests, analyses
active seismic earth pressure, passive seismic earth pressure, for slope stability or deformation are conducted. In the analyses
excess pore water pressure and dynamic water pressure. Using of slope stability during earthquakes, seismic force and excess
external forces, a stability analysis was conducted that included pore water pressure must be considered as shown in Figure
sliding, overturning, bearing capacity of the stratum, and 4.16. Then, in the guideline, methods to evaluate excess pore
circular slip failure. The detailed design method of Arakawa water pressure are introduced at first. One simple method to
Dike is explained in Yasuda (2007). estimate excess pore-water pressure due to liquefaction the is
same as mentioned above in 4.2.4 setting Ru to 1.0 for FL≤1.0
(Japan Road Association, 1986).

Figure 4.15: Remediation methods for existing embankments

4.5.6 Introduction of seismic design for reclaimed area for Figure 4.16: Time of the occurrence of seismic force and excess pore
water pressure
housing lots
Though slope failures in housing lots cause severe damage to
As illustrated in Figure 4.16, little excess pore water pressure
houses and loss of lives, much reclaimed land has been
is usually induced before the peak of seismic motion (Ishihara
developed in urban areas without considering seismic stability.
and Yasuda 1975). Therefore two sets of analyses (a) to
Then appropriate procedures are needed to evaluate seismic
consider seismic coefficient only, and (b) to consider excess
stability of existing reclaimed areas. Recently, this procedure
pore water pressure only, are recommended for the evaluation
has been discussed and a guideline has been proposed by the
of seismic slope stability, as shown in Table 4.10.
Kanto Branch of the Japanese Geotechnical Society (2007), see
In performance-based design, it is necessary to estimate not
also Yasuda (2007). In the guideline, it is recommended to
only the stability but also deformation of embankments. Several
evaluate the seismic stability by the following five steps.
methods to estimate the deformation of embankments are
Step 1: Finding out of filled zone
introduced in the guideline. These are:
In a reclaimed area for a housing lot, it is difficult to find the
1. Dynamic response analysis
boundary of cutting and filling by site survey, if there is no
2. Newmark’s method
record on reclamation work. Then the guideline recommends
3. Static residual deformation method
finding the filled zoned by comparing old and new aerial
Step5: Selection of appropriate countermeasures
photos. The filled zone can also be estimated by comparing two
The final step is the selection of appropriate countermeasures
sets of topographical maps drawn before and after the
against slope instability during earthquakes. Three types of
reclamation work. However the accuracy of the estimated zone
countermeasures against instability of existing fill slopes are
is likely to be lower than that measured by aerial photos.
introduced in the guideline: control works, preventing works
Step 2: Field investigation to select unstable slopes
and slope protection works. Some of these are listed in Table
The next step is the identification of existing fill slopes
4.11. Examples of typical arrangements are shown in Figures
which will become unstable during earthquakes. This can be
4.17 and 4.18.
evaluated roughly by visual inspection, based on several factors
shown in Table 4.8.
2902 B. Simpson et al. / State of the art Report: Analysis and Design

Table 4.9: Rough check sheet by visual inspection for existing retaining wall prepared by Yokohama City.

Value
Stone wall RC wall
Classific- Item for check Normal Slightly Abnormal Normal Slightly Abnormal
ation abnormal abnormal
Drainage Drain hole 0 1.0 2.0 0 1.0 2.0
condition Ground behind wall 0 1.0 2.0 0 1.0 2.0
Percolate from wall 0 0.5 1.0 0 0.5 1.0
Inclination of the ground behind 0 1.0 2.0 0 1.0 2.0
wall
Drainage facilities of the ground 0 1.0 2.0 0 1.0 2.0
behind wall
Wall Height and inclination of wall 0 2.0 4.0 - - -
Horizontal crack of wall 0 4.0 6.5 0 3.0 5.5
Vertical or cross crack of wall 0 2.5 5.0 0 1.5 4.0
Crack at the corner of wall 0 3.0 5.5 0 2.0 4.5
Horizontal movement 0 3.5 6.0 0 2.5 5.0
Differential settlement 0 4.5 7.0 0 3.5 6.0
Opening of the corner of wall 0 4.5 7.0 0 3.5 6.0
Expansion of wall 0 5.0 8.0 - - -
Tilt of wall 0 5.5 9.0 0 4.5 8.0
Corrosion of iron bar - - - 0 5.0 8.0
Total value
Judge Total value is less than 5 Judge as less abnormal by inspection
Total value is greater than 5 and less than 9 Judge as abnormal by inspection
Total value is greater than 9 Judge as quite abnormal by inspection

Table 4.11:Countermeasures against instability of existing fill slope and


arrangement plans
Type of Countermeasure methods
countermeasure
Control works Surface water drainage work, Groundwater
drainage work, Ground water cut-off wall, Soil
removal work, Counterweight fill method
Preventing works Retaining wall, Soil reinforcement, Ground
anchorage, Pile works, Shaft work
Slope protection Turfing, Grating crib work, Gunite, Sprayed
works concrete, Gabion

4.6 Recent developments - cut slopes


Figure 4.17: Countermeasures by horizontal drainage drilling and
drainage wells (Kanto Branch, JGS, 2007) During heavy rainfalls it is necessary to control the traffic of
cars or trains. Rain gauges are effective for judging the critical
condition to shut down the traffic. As a result, many rain gauges
have been installed and many methods to judge critical
conditions have been proposed. In the proposed methods
several factors are considered (see, for example, ATC3 1997) :
1. Total rainfall (for example, two weeks up to the previous
day)
2. Power of rainfall
3. Rainfall in one hour
4 Effective rainfall
5. Tank model method
Figure 4.19 shows one diagram to judge critical conditions
by continuous rainfall and rainfall in one hour. Boundaries
between failed and safe cut slopes and embankments estimated
Figure 4.18: Countermeasures by ground anchors and pile work (Kanto
Branch, JGS, 2007)
based on case histories are shown by curves. The limit to shut
down the Japanese railways is shown by step lines.

Table 4.10: Two factors to be considered in stability analysis. 4.7 Examples of failures- embankments

Factors to be considered
Type of filled soil Excess pore Seismic 4.7.1 Slope failures and slumps of expressway embankments
water pressure coefficient during the 2004 Niigataken-chuetsu earthquake in
Loose and yes No Japan (partially quoted from Yasuda et al. 2008)
saturated On October 23 in 2004, the Niigataken-chuetsu earthquake, of
Sand
Dense and/or No yes Magnitude 6.8, occurred and caused serious damage to many
unsaturated structures and slopes in Japan. Six expressways were closed due
Clay no yes
B. Simpson et al. / State of the art Report: Analysis and Design 2903

Several seismic records were obtained in these severely


damaged zones. The recorded maximum surface accelerations
were 0.500g at Horinouchi Town, 1.757g at Kawaguchi Station,
1.532g at k-net Ojiya site and 1.029g at Ojiya Castle. Therefore
it can be said that seismic motion in the severely damaged zones
was very strong as the maximum surface acceleration was about
0.5g to 1.7g.
A section between Koide IC and Ojiya IC of Kan-etsu
Expressway was selected to study the influence of type of
embankments upon the damage. In this section, embankments
were constructed by three methods: filling on a level ground,
widening, and half-bank and half-cut. Percentages of the lengths
constructed by these methods are 57 %, 5 % and 7 %,
respectively. The other 31 % are cuttings, tunnels and bridges.
Total lengths of damaged and undamaged embankments
constructed by the three methods are compared in Figure 4.21.
Lengths of damaged embankments seem to be two to three
times the length of intact embankments regardless of
construction method, and many sites were damaged even where
Figure 4.19: Diagram to judge critical conditions of failure of cut slopes the ground is flat.
and embankments for Japanese railways In Japan, damage to road embankments is classified in three
levels as shown in Figure 4.22, and this was applied to the Kan-
to the earthquake. The total length of the closed expressways etsu expressway. Serious damage occurred at half-bank and
was 580 km. Emergency treatments were applied to the half-cut sections only. In the embankments on level ground,
damaged expressway embankments by filling, placing and medium or minor damage dominated. According to the
spreading. Then all expressways were able to opened for mechanism of failure, the damage of the Kan-etsu expressway
emergency vehicles about 19 hours after the earthquake because embankments in the section between Koide IC and Ojiya IC,
no serious damage was induced in expressway bridges and can be classified to three types as follows:
tunnels. About 13 days after the earthquake all expressways Type 1: Serious slide of the embankment on the sloping
were opened for all vehicles. ground as schematically shown in Figure 4.23(a)
Among the affected six expressways, the following two Type 2: Settlement of the embankment on the level ground
zones were severely damaged. without the deformation of the ground as schematically shown
1. Between Muikamachi IC and Nagaoka IC of Kan-etsu in Figure 4.22(b)
Expressway (57.6 km) as shown in Figure 4.20, and Type3: Settlement of the embankment and the culvert on the
2. Between Kashiwazaki IC and Sanjyo-Tsubame IC of level ground with the deformation of the ground as
Hokuriku Expressway (50.3 km) schematically shown in Figure 4.223c)
Most serious damage occurred in the following sections: Locations where these types of failures occurred are shown
a. Kan-etsu Expressway: between Horinouchi IC and on Figure 4.20.
Echigokawaguchi IC (8.8 km), and between
Yamamotoyama Tunnel and Yamaya PA (5.5 km ) ,
b. Hokuriku Expressway : between Ohzumi PA and Ngaoka
JCT (6.0 km)
The section between Horinouchi IC and Echigokawaguchi
IC of Kan-etsu Expressway was constructed on the gentle
slopes of hills. In contrast, the section between Yamamotoyama
Tunnel and Yamaya PA was constructed on flat ground. In the
former section, embankments were constructed mainly by “half-
bank and half-cut” methods on the slope of the hills. Sliding of
the filled embankments occurred at several sites during the
earthquake. Where embankments were constructed by filling
soils on level grounds in the latter section, large settlements of
the embankments occurred.

Figure 4.21: Total length of damaged and not damaged embankments


constructed by three methods

In Japan, the current seismic design method for road, railway


and river embankments is the seismic coefficient method based
on circular slip surface analysis. However, it has become
necessary to introduce performance-based design, in which it is
necessary to evaluate the performance by the deformation of
embankments, such as settlement. However, analytical methods
to evaluate the deformation during earthquakes have not been
fully developed. So, based on these case histories for road,
railway and river embankments, new design methods have been
discussed by several technical committees organized in the
Figure 4.20: Route map of Kan-etsu Expressway and Hokuriku Japanese Geotechnical Society or other associations. In a few
Expressway
2904 B. Simpson et al. / State of the art Report: Analysis and Design

years, new seismic design methods will be introduced in design soil liquefied during the earthquake and caused settlement and
codes for road, railway and river embankments. In the new tilting of houses.
design codes, the following analytical methods will be In 2003, 35 years after the 1968 Tokachi-oki earthquake, a
introduced: new Tokachi-oki earthquake occurred and caused damage to
Type 1 – Embankment on sloping ground: Newmark’s method. houses in this district again. Figure 4.25 compares damaged
Type 2 – Embankment on level ground: Dynamic response houses during the two earthquakes. In both cases, damaged
analysis or static residual deformation analysis. houses were located on fill ground. However, houses damaged
during the 1968 earthquake survived during the 2003
earthquake. After the 1968 earthquake underground conduits
were constructed in the damaged zone, probably lowering the
ground water level. This may be one of the reasons why the
damaged houses survived during the 2003 earthquake.

Figure 4.22: Classification of damage to road embankments

Figure 4.24: Soil cross section and damaged houses during the 1968
Tokachi-oki earthquake

At Utsukushigaoka district near Kiyota district, several


houses settled as shown in Figure 4.26 during the 2003
Tokachi-oki earthquake. It can be judged that liquefaction
occurred in this site because traces of sand volcanoes were
observed around the damaged houses. The boiled soil was
volcanic sandy silt. The damaged ground had been constructed
recently by filling a channel in a hill zone. Therefore, it is
considered that the filled soil liquefied and caused the
settlement of the houses. Yasuda et al. (2004) measured the
angle of inclination of the damaged houses roughly, as indicated
on Figure 4.27. As shown in this figure, about seven houses
settled and tilted more than 1°. According to the experience of
the damage during the 2000 Tottoriken-seibu earthquake,
Figure 4.23: Classification of the damage to the embankment of inhabitants felt giddy and nauseous, and after the earthquake
Kan-etsu Expressway according to the mechanism of failure. they could not live in the houses which tilted more than 1/100,
Type 1: Serious slide of the embankment on the sloping ground; as indicated in Figure 4.26
Type 2: Settlement of the embankment on the level ground without the
deformation of the ground;
Type3: Settlement of the embankment and the culvert on the level
ground with the deformation of the ground.

4.7.2 Slope failures of a reclaimed area for housing lots


1. Liquefaction-induced settlements of houses on filled
grounds
An artificially filled housing lot at Kiyota district in Sapporo
City was severely damaged during the 1968 Tokachi-oki
earthquake. This might be the first experience of the damage to
artificially filled housing lots due to earthquakes, in Japan. The
maximum surface acceleration at Kiyota was estimated as about
0.08 g. The housing lot was constructed by cutting soils from
hills and filling valleys as shown in Figure 4.24. There were
279 houses in the housing lot. Of them, 27.5% of houses were
damaged during the earthquake. As shown in Figure 4.24 no
house constructed on the cut ground was damaged. On the
contrary, 56% of houses constructed on the filled ground settled
and tilted. Ground water table in the fill was shallow and filled Figure 4.25: Comparison of damaged houses during two earthquakes
soil is volcanic ash sand. Therefore it is considered that the fill
B. Simpson et al. / State of the art Report: Analysis and Design 2905

Figure 4.26: Settled houses due to liquefaction during the 2003 Figure 4.28: Collapsed houses due to slope failure at Midorigaoka in
Tokachi-oki earthquake at Utsukushigaoka district Sendai City during the 1978 Miyagiken-oki earthquake

Figure 4.29: Procedure of reclamation work conducted at Midorigaoka

Figure 4.30: Soil cross section along failed slope (Tohoku Branch,
JSCE)

Figure 4.27: Rough measurement of inclination of settled houses

2. Collapse of houses due to failure of fill slopes


Several housing lots were damaged during the 1978
Miyagiken-oki earthquake in Sendai City in Japan, some of
them affected by a severe slope failure of an embankment at
Midorigaoka Ichi-chome housing lot. As shown in Figure 4.28,
many houses collapsed due to the slope failure. In 1957 to 1958,
the embankment was constructed by filling a valley. During the
reclamation work, trees on surrounding hills were felled and
thrown down into the valley. Rocks cut from the hill also were
thrown into the valley, then soil cut from the hills was placed as
schematically shown in Figure 4.29. Figure 4.230 shows a soil
cross-section along the slope. The thickness of the fill was 5 to
15 m. The density of the fill was very low with 0 to 10 of SPT
N-values. It is estimated that buried trees became humus within Figure 4.31: Failed slope and damaged houses at Midorigakoka in
Kushiro City during the 1993 Kushiro-oki earthquake
30 years. Moreover, as big voids existed between buried rocks,
groundwater flowed in the big voids and caused weathering of
At Kayanuma district in Shibecha Town, land for housing
rocks. Then the bottom of the fill became very loose and caused
lots had been developed by cutting the north slopes and filling
the slope failure during the Miyagiken-oki earthquake.
in the south zones as shown in Figure 4.34. Four valleys were
During restoration, steel pipe piles, 318.5 mm in diameter,
filled with a maximum thickness of 10 m. No special drainage
were installed to increase resistance against sliding, as shown in
conduits were put on the bottom of the valleys. Filled soils slid
Figure 4.30. The steel pipes was strengthened by filling with
and nine houses were severely damaged during the 1993
steel H sections and concrete. The steel pipe piles were installed
Kushiro-oki earthquake as shown in Figure 4.35. Soil
in two rows at 2m centres. A concrete retaining wall and
investigations by Swedish Weight Sounding and other tests
drainage wells were also constructed. The safety factor against
were carried out after the earthquake to study the mechanism of
sliding during anticipated earthquakes was thereby increased to
damage. The fill soil was volcanic silty sand. Figure 4.36 shows
1.2.
a soil cross section along lines A-A’ and B-B’ which cross
heavily damaged zones. The filled soil was loose with SPT N-
2906 B. Simpson et al. / State of the art Report: Analysis and Design

vales of around 10, and the lower part of the fill was saturated.
It was difficult to recognize the occurrence of liquefaction
during the site survey conducted just a few days after the
earthquake, because the land was covered with snow and sand
volcanoes could not be observed. However, detailed seismic and
liquefaction analyses suggested the occurrence of liquefaction
of filled sand. In the slope stability analysis, the safety factor
against sliding was less than 1.0 if the seismic acceleration
exceeds 0.2g Unfortunately the actual acceleration is not known
because no seismic records were obtained around here.
Deformation analysis by Tara 3 showed large deformation of
about 1m, as shown in Figure 4.37. Deformation of the area
could be estimated by FEM as shown in this figure because the
slide was not large.

Figure 4.35: Deformed bank and damaged houses at Kayanuma during


the 1993 Kushiro-oki earthquake

Figure 4.36: Soil cross section along the slope (JSSMFE, 1994)

Figure 4.32: Location of damaged houses at Midorigaoka


(JSSMFE,1994)

Figure 4.37: Deformation of bank analysed by Tara 3 (JSSMFE, 1994)

5 UNDERGROUND CONSTRUCTION

5.1 Introduction

This section of the paper addresses design and construction of


underground structures, principally tunnels, and presents
observations and findings based on experience gained over the
Figure 4.33: Mechanism of slope failure (JSSMFE, 1994) past two decades during the design and construction of major
railway, highway and water-transfer tunnel projects in the
United States, Africa and Asia. During this time it has been
possible to compare the performance of the as-constructed
structures to the predicted performance as given by the design
process. Excavation was typically carried out in widely varying
ground conditions (weak-bedded rock, jointed hard rock, and
soil-like poorly cemented rock with groundwater). It is therefore
intended that the design guidelines could be applied, in general
terms at least, to any tunnel excavated in any ground mass
condition with any ground material type (hard rock, soft rock,
bedded rock, soft ground, and soil) with or without ground
water.
It may be queried as to why there is a basic review of tunnel
design in a state-of-the-art paper. The purpose of this
rudimentary review is to have designers and other design team
members, inclusive of managers, realize that even with the
advances of numerical methods and the tools available to assist
in technical evaluations, it is still essential not to lose sight of
the fundamentals. Common sense and a basic understanding of
Figure 4.34: Filled zones and damaged houses at Kayanuma (JSSMFE,
1994)
the fundamental physics involved, along with comprehensive
checks during both design and construction are still imperative,
even in the presence of all the computer wizardry.
B. Simpson et al. / State of the art Report: Analysis and Design 2907

The paper provides a basis for design of underground would also be required for contractual, insurance purposes, etc,
structures, specifically tunnels, by separately addressing the in the event of unsatisfactory performance of the structure.
following: Basis of Design, Design Methodology, Design Codes During the design process typical Design Management
and Analysis, Materials and Sustainability, Performance Elements include but are not limited to:
Monitoring and Construction Reviews and Future • Collect data and develop concept
Developments. • Project hazard analysis
Increasingly, concern for the environment and spiralling • Defining design requirements
costs for commodities such as steel, concrete and fuel have also • Define design protocol of codes and standards.
driven owners to consider and introduce specific requirements • Define quality processes
for sustainability. So one intention of this paper is to • Technical risk management
demonstrate how these sometimes conflicting requirements can • Technical studies
be addressed through innovation and quality in the design, • Interface management
procurement and construction process. • Technical review
• Environmental management
5.2 Basis for design • Internal value engineering
• Producing designs and documents
• Clarifications
5.2.1 Definition of design • Correspondence, mtgs. and communications
The basics of tunnel design and achieving the functional • Change identification and implementation
requirements through the design of large span soft ground Careful management of these issues will assist in the
tunnels based on the sequential excavation and support method development of an appropriately documented integrated design
(SES) or Cut-Cover tunnels on a design-build or design-bid- as is discussed later in this section.
build contract delivery project requires a well planned process
that defines the objectives of the design as well as the process to 5.2.2 Types of underground structures
follow while executing the design. This first essential For the purpose of this paper, a tunnel is regarded as an
requirement is usually addressed through the contract underground opening of limited span (width) and height
documents which govern the administration of any given (generally less than about 15 m), oriented in a horizontal/sub-
project. horizontal direction, which is advanced linearly by rapid
As more projects follow the trend of transferring project excavation techniques generally through a variety of geological
design and construction to a single entity and design service life materials. This section will not specifically cover other
increases to 80, 100 or 120 years, it becomes more important underground excavations such as shafts (which are oriented in a
that contract documents clearly state the requirements expected vertical/sub-vertical direction) and caverns, which have
of the design. Continuity in design process becomes paramount significant span and height and are more three-dimensional in
in ensuring that the finished product meets the expectations of nature.
the Owner. This is clearly evident for design-build contracts, Cut-and-cover construction, as the name implies, involves
but often the traditional design-bid-build projects also suffer excavating a trench (cut), constructing the tunnel structure, and
from a lack of proper understanding of the development of backfilling to restore the surface to its original condition
tunnel design. (cover).
A tunnel design typically consists of six major attributes The concept of cut-and-cover construction dates back to at
which define the required work: least 2180BC, when Babylonians constructed a brick arched
• Each tunnel design is individual tunnel, approximately 12 ft. high and 15 ft. wide in a trench
• Each tunnel design requires the excavation method to be across the Euphrates River as noted in the Questia Online
selected Encyclopaedia (www.questia.com/library/encyclopedia/tunnel.jsp).
• Each tunnel requires the decision to be either Drained or The river flow was first diverted prior to construction and
Undrained which defines the type of loading the lining must reinstated after the tunnel was backfilled.
withstand With its inevitable surface disruption, cut-and-cover
• Each tunnel design should define the standards and construction in highly trafficked urban environments can be
analytical tools to be used intrusive. However, for a number of reasons – shallow
• Each tunnel design typically has a unique set of “Owner” alignment, large and/or varying cross section, prevailing ground
requirements conditions – which may make conventional tunnelling risky
• Each tunnel design is not complete until the excavation is and/or uneconomic, cut-and-cover construction remains a viable
completed and the correct lining installed. and popular alternative.
It may be asked, what is design? It may be thought of as any Bored tunnels are typically employed when costs of cut-
rational/verifiable/justifiable process whereby a concept is cover tunnels exceed that of mining techniques or surface
transformed into information that can be used to bring the disruptions are not tolerable. Tunnel bores may be advanced by
concept to reality. In the case of a civil structure, design may be a tunnel boring machine (TBM), generally with the use of a
defined as that process, which is documented and can be segmental concrete or iron lining, or by other techniques.
audited, that determines information such as the shape (form) of Approaches which do not use a TBM include SES –
the structure, the dimensions (size) of the structure, the sequential excavation and support method, also known as SEM
thickness and strength of structural members that comprise the – sequential excavation method. This form of excavation
structure, etc. This information can be used to compute the requires providing support in the form of ribs and lagging or
time, money and resources required for construction of the shotcrete and so is also known as sprayed concrete construction.
structure. An optimum design may be thought of as one that not In the highly stressed rocks of the European Alps, this approach
only maximizes the functionality and safety of the structure and has been developed as the NATM – New Austrian Tunnelling
minimizes its cost, both from an initial (capital) cost and a long- Method – in which support is provided to the opening based on
term (maintenance) point of view, but also takes advantage of the observed behaviour of the ground while adjusting to the
known sustainability options available as well. revised stress fields.
A very important requirement of a sound design is that there
exists documentation of the process that was used, so that the 5.2.3 Performance requirements
design work can be examined later in relation to the measured Owners’ performance requirements for underground structures
or observed performance of the structure. Such an audit review have traditionally included safety, economy and durability for
2908 B. Simpson et al. / State of the art Report: Analysis and Design

both construction and subsequent operation over and beyond the • Serviceability requirements (allowable
specified design life of the structure. displacement/deformation/cracking)
To assist with a common understanding of design terms, the • Environmental issues / handling of tunnel spoil /
following definitions are suggested. However any given project groundwater pumping
often has a set of predefined conditions to which the design • Project cost restraints / budgets / schedule
must adhere. It is suggested that when this occurs the project However, even with properly defined Owner Requirements
team clearly define the terms and how they are intended to be there must always be a clear traceable path of evidence that
used in the design of the structure. Following this suggestion shows that the client requirements for function, operation,
will help to eliminate any misunderstanding between Owners, finished size and shape have been achieved and that life-safety
Designers and Contractors. issues have been identified and duly addressed.
Tunnels must be safe as society demands security in the The Role of Geotechnics
structures it inhabits or uses – thus tunnels are often classified An important difference between tunnel design and any other
into two categories: civil construction is the percentage that geotechnical parameters
• Inhabited space influence the design and construction and are quantifiable in
• Uninhabited space comparison to the other design data. For example, in the design
These definitions help clarify the level of risk that needs to of a bridge or building, as illustrated in Figure 5.1, the
be managed and where that risk may reside, i.e., life safety or geotechnical variables influence only the foundation, which is a
project capital costs. minor part of the construction, and so give rise to little financial
Having defined the intended use of the underground space, it risk during construction. Whereas in tunnel design, these
is imperative to understand the requirements for which the parameters influence the excavation and support which is the
structure is required to perform. The definitions of the types of greater part of the design and construction; this results in a high
analyses to be incorporated in the design provides a sound basis risk on a major part of the financial cost associated with the
for guiding the design process. The definitions below serve well work.
to guide the designer in this effort and are closely related to the Hence the age old saying:
functional requirements that are discussed below.
ULS – Ultimate Limit State Design may be viewed such that “Geotechnical Investigations?? ….. Pay Now or Pay Later”.
exceeding the limit results in failure. It is irreversible and may
result in risk to loss of life and/or structure Accordingly, the value of good geotechnical data and
SLS – Serviceability Limit State Design may be viewed as a applied geotechnical engineering is essential to a successful
hindrance or undesirable performance, but this may be tunnel project.
reversible and does not risk loss of life or structure.

5.2.4 Functionality requirements


The type of tunnel to be designed and constructed often defines
the requirements. Generally there are five types of tunnels and
each category generally has its own specific requirements for
functionality. The most common types of civil works tunnels
are:
• Road Tunnels
• Railway Tunnels (Mass Transit, Commuter, Freight, High
Speed)
• Hydraulic Tunnels (Water, Wastewater, Storm Water,
Head/Tail Race)
• Utility Tunnels
• Pedestrian Tunnels
Each of the above tunnels generally requires a clear
statement of performance requirements regarding but not
limited to the following: Figure 5.1 Role of geotechnical influence on design
• Tunnel cross-section(s) (twin bores versus single bores)
• Design life 5.2.5 Site investigations
• Tunnel orientation A tunnel is a structure that is typically controlled
• Durability geotechnically. The predominant loads acting on the structure
• Alignment (horizontal and vertical)
are the ground pressure load and, if groundwater is present, the
• Lighting groundwater pressure. Other loads include dead load associated
• Portal locations with the liner and any non-removable elements and the live
• Power
loads as well as surcharge loads.
• Structure gauge/size A tunnel design is only as good as the geological and
• Aerodynamics geotechnical data. An important step in the design process is the
• Water tightness /leakage
determination of a “best estimate” of the anticipated geological
• Noise and geotechnical conditions along the planned tunnel alignment.
• Ventilation/purging This information should include the following:
• Vibration
• Ground types (e.g. sandstone, silt, clay, sand, etc)
• Drainage • Geological structure (bedding, jointing, faults, shear zones,
Other issues which generally require early decisions include: etc)
• Fife safety issues and requirements
• Ground mass strength and stiffness
• Security requirements • Groundwater characteristics (location of water table(s),
• Orientation of tunnel with regards to geological setting / permeability, etc).
design requirements
An adequate site investigation program prior to construction
• Drained or undrained tunnel design (usually should always be carried out for this purpose. The work must
environmentally driven) be sufficient to identify any adverse conditions which might
B. Simpson et al. / State of the art Report: Analysis and Design 2909

occur along the tunnel alignment. Inadequate data often leads regardless of the type of hydrostatic loading envelop applied.
to excavation delays, slower progress, which increase the time However, for non-circular tunnels, the tunnel invert will require
required, and thus the cost, to excavate, support and line the significant reinforcement depending on the type of hydrostatic
tunnel. envelop applied and the geometry of the tunnel section.
It is often not possible to determine the exact distribution of Drained tunnels are typically required to have longitudinal
conditions along the tunnel drive, so often “representative” groundwater drain pipes along each side of the tunnel with
conditions are established. These then serve as “benchmark” maintenance access not to exceed 50 m intervals. Where drained
conditions that can be later compared to the actual conditions tunnels are constructed the drains typically consist of perforated
encountered. Historically, for any given region where pipe with either a gravel pack or no fines concrete for collection
geological conditions are equivalent and reasonably understood, of the groundwater. The pipes are located near, but always
it is possible to arrive at an estimated percentage distribution of below, the longitudinal construction joint at the interface
ground types to be assumed for use in the “base” design for between the crown lining and the invert ‘corner’. The inner
bidding. The actual design will require verification of ground lining is tapered above and below to provide space for the
conditions during construction with appropriate payment terms. drainage collection pipes.
Once a prediction of representative geological and Undrained tunnels are subjected to the full hydrostatic load
geotechnical conditions along the planned tunnel drive has been as defined by the design groundwater level for the project or
achieved, the next step is “geotechnical design”. This is the section thereof. Some projects establish more than one design
process of predicting the response of the ground mass in each groundwater level. In all cases save transient load cases,
representative condition along the tunnel to the planned appropriate load factors should be applied to the groundwater
excavation of the tunnel. This prediction is required to assess pressures when designing the tunnel unless a “working stress”
what construction method, excavation method, support analysis is performed. Differing design procedures apply the
measures and ground treatments may be required to achieve load factors either to water pressure data in the analysis or to
stabilization of the excavation. For small span tunnels in “good” resulting stresses in the linings. Further debate of the merits of
ground minimal, if any, measures are required and routine these alternatives will be valuable, involving both tunnelling
construction and excavation methods are possible. In “bad” and geotechnical disciplines. A very applicable paper by
ground, however, particularly where the tunnel is large in size, Bilfinger (2005) provides an approach to designing for
extensive complex procedures may be required. groundwater loads on tunnel linings.
It is also relevant to note that tunnels do not have to be dry in
5.2.6 The role of water and water pressure order to achieve an undrained pressure loading. The only
Of particular note, the designation of a drained or undrained condition that is required to attain full hydrostatic load to the
tunnel is not a definition of water tightness but it is a clear tunnel lining is a low relative permeability of the tunnel lining
definition of the intended load which the tunnel shall be relative to that of the ground surrounding the tunnel.
designed to resist. A clear understanding of this design
requirement is essential to avoid catastrophic errors in the 5.2.7 Waterproofing
selection of the excavation opening size and the loading which The Owner’s requirements will usually dictate the required
the permanent liner shall be designed to resist. Undrained “dryness” of a tunnel. Water tightness usually has a direct
tunnels are typically designed to resist the full water pressure as impact on operation and maintenance costs for the Owner if the
defined by the designated groundwater table and geo- underground space is considered as inhabited. Accordingly,
hydrological model. Whereas drained tunnels are often designed water tightness should be specified by an allowable leakage
to resist only some portion (usually minimal/residual) of the rate. It is rarely observed that a tunnel is truly dry. Hence there
pre-existing hydrostatic pressure. should be an allowance for acceptable leakage. This allowance
The decision to build an undrained or drained tunnel should requires an easy but reasonably discernable method to
be based on actual requirements. Environmental issues should determine contract conformance.
be the most significant factor contributing to the selection of an During the course of a recent project it became clear that the
undrained tunnel. For example, it is understood that the German contract documents were incomplete in addressing the
Rail preference has recently followed this policy of making this requirements for water tightness for a rail tunnel where the
selection rather than the previous requirement for design and contract merely stated that the tunnel must be “dry”. Often
construction of only Undrained tunnels to reduce structure complete dryness is not required to meet the functional
maintenance. requirements of the structure or that of the Owner. Accordingly
Often the Designers are permitted to propose design and the subject was duly reviewed and further clarified to provide a
construction of drained tunnels where they consider the design practical measure to achieve the Owner’s functional
appropriate unless they are specifically prohibited by Contract. requirement. To meet the intent of the project Design
The following conditions typically need to be satisfied to allow Specifications which stated that the tunnels shall be constructed
for the design and construction of a drained tunnel: to be completely “dry”, it has been recommended that the use of
• The reduction of the water table will not have adverse the German Code DS 853 Class 3 as the objective criterion for
environmental effects (such as settlement of structures, evaluating compliance with this requirement. Unless very
significantly reduced stream flows, etc.). stringent watertightness criteria are required as for underground
• The inflows would not be so high as to cause maintenance system control facilities, etc., this class of “watertightness” is
problems. considered adequate for most inhabited underground space as
• The ground water is not aggressive (<1.5 g/l of sulphates). typically required for transit projects.
Drained tunnels are typically required to resist some amount Typical water proofing systems consist of geofabrics,
of residual hydrostatic pressure. Often this pressure envelop synthetic geomembranes with a series of water stops to
consists of an allowance for a residual radial water head of 5m compartmentalize segments of the tunnel structure and post
above the inner contour line of the inner lining. However, a grouting tubes to seal off leakage in the future. Designs
proper understanding of the hydrogeological characteristics of typically include a waterproof membrane (typically 2.5 mm
the geology where the tunnel is driven is required. In some pvc) over the crown. A geotextile ‘fleece’ is usually provided
cases the residual radial pressure envelop approach is outside the membrane and external waterstops cover the full
appropriate. In other cases a more conservative linearly perimeter of all transverse construction joints. Isolating
increasing hydrostatic pressure loading may be more waterstops are generally required at bulkhead locations to
appropriate. The tunnel arch or vault area usually is not prevent water leakage from extending beyond the space
significantly affected by this amount of hydrostatic load between the bulkheads.
2910 B. Simpson et al. / State of the art Report: Analysis and Design

Where undrained tunnels are required or proposed, • Feasibility Study – choice whether to construct a tunnel or
Designers typically provide a full perimeter waterproof not
membrane (typically 2.5 mm pvc) to create a fully “tanked” • Conceptual Design – choice of type of tunnel and alignment
structure. • Preliminary Design – selection of initial design options,
such as overall geometry and material grades
5.3 Design methodology • Detailed Design – selection of final design options
• Construction Design – selection of construction options and
As with all project designs, a clear understanding of the design as-builts.
methodology must exist which is geared to addressing all the For each level of the design process, the project begins with
Owner’s requirements and those issues not directly addressed an idea and each subsequent phase of the design advances and
by the Owner but which are the responsibility of the Designer to requires more detail to be developed. With the completion of
deliver a durable and safe system as defined by the Designer’s each phase of the project various elements of the design are
profession. established and design options assessed and presented which
allow the “decision-maker” to determine the next course of
5.3.1 Integrated tunnel design process action for the subsequent phases of the project. There are
Tunnels must be economical firstly with regard to cost to design always elements of the design which must be worked out in
and construct and secondly with regard to maintenance. Most detail in the earlier phases of the project, which then will be
design decisions are, implicitly or explicitly, economic required to be properly assessed to evaluate the effects and
decisions. The life-cycle evaluation of the structure accounting resolve issues that will be encountered in later phases of the
for all costs and benefits arising from all the phases of the life of project.
the tunnel must be taken into account. These phases include
design, construction, operation and closure or transformation 5.3.2 Design checks
(salvage) of the facility. This concept is rapidly being advanced The design check list given below has attempted to show the
in the European countries under the designation DARTS hierarchy of the design steps which are required to complete the
(Rostam & Høj 2004) which is an acronym for “Durable and actual design of the structure once the Owner’s requirements
Reliable Tunnel Structures”. and functionality issues have been resolved.
To account properly for the life cycle costs of the structure, • List functional requirements
an integrated design should account for all effects on the Owner • Select appropriate design codes of practice and design code
and society over the complete lifetime of the structure as an protocols
integrated process. The level of detail requires adjustment to • Determine design loading (empirical or numerical methods)
match the current phase of the planning and design and • Select applied design service loads / load factors / load
subsequent decisions, which affect the final structure and are combinations
required at various stages of the planning and design process. • Perform structural analysis
This approach requires that all parties involved with the project • Establish required strength / internal forces / moments /
cooperate as a team and not as contractual opponents. An stresses
adversarial relationship during a project is a recipe for • Select material strength and reduction factors based on
marginalizing the design, and risks missing the optimization of strength requirements
the project. The product, as seen by the Owner and Designer, • Size/re-size structural elements and determine design
will influence the design, specification, contracts, and the strength capacity
organization and structure between the various parties during • Check design strength capacity against required strength
execution and operation. Obviously this affects how the demand
interfaces between the Owner, Designer, Contractor and other • Determine serviceability requirements
third parties are selected and administered over the life of the • Material stress (as applicable)
project to result in an integrated design. This is important as • Crack control
most tunnels are designed and constructed to serve the public • Deflections/deformation
and the complete life-cycle of the project should be optimized to • Finalize design based on constructability and material
their benefit availability
During the process of modelling and selecting design options • Check, check and recheck the calculations and assumptions
and solutions, there is always uncertainty in the quantification Although the above may appear to be common sense, it is
each issue. All potential outcomes of the design process must be surprising how often errors in the detail of the design have
considered, evaluated and documented. Omission of any resulted in large design, construction and contractual problems
potential outcome may incorrectly favour solutions which do for a project. Peer reviews or group review sessions are always
not provide real benefit. very helpful in identifying shortfalls in any design. Designers
Uncertainty may not be limited to the physical or statistical and various team members should learn that these reviews are
sense but also includes the relative belief in that uncertainty geared to identify potential short falls or problems and
developing over the life of the structure. Examples of this sort determine the most appropriate resolution. The goal is to deliver
of item include but are not limited to: 1) reliability of the the most optimal design for the project.
structure performance with respect to limit state analyses, With today’s computers, often the designers fail to grasp
structural behaviour, durability, etc., 2) uncertainty of design how the analytical tools currently available for use imply far
loads developing and particularly in regard to load conditions or greater precision than our ability to understand the ground
combinations, 3) uncertainty of costs associated with design properties for which the design must account.
modifications, geotechnical conditions, and environmental Sensitivity analyses and an ability to understand the potential
impacts, and 4) probability of undesirable events. hazards and risks associated for the design can help identify
The design process may be viewed in the feasibility phase where real savings in the design or real risks in the development
having decisions revolving around financial and environmental of undesirable performance may exist. Often the sensitivity
issues (politicians, bankers, government) whereas the detailed analyses will assist in identifying if additional design effort is
design and construction phases are focused on the technical really beneficial to the project or not.
engineering and constructability (future Owner/Operator and In the past, inappropriate decisions in the application of
Contractor) aspects of the project. codes and design methods have led to extraordinary events
Typically the project design is divided into five phases; which may render a structure useless or far worse create a
potential for injury or loss of life. A recent example is the
B. Simpson et al. / State of the art Report: Analysis and Design 2911

collapse of the Nicoll Highway excavation in Singapore between itself and the diaphragm wall. This gap allows
(Simpson et al 2008). This issue returns the reader to the sufficient space for a camera scope or similar remotely operated
beginning of this section where it is essential that the Owner, device to be inserted for inspection purposes.
Designer and Contractor clearly understand the functional Similar issues surround that of tunnel drain inspection and
requirements of the structure and the performance limits that the maintenance of drained tunnels. This inspection and
structure demands. maintenance requirement is critical to tunnel performance as the
Issues regarding proper modelling of construction sequences drains control the magnitude of load the tunnel structure has to
and load development, load combinations, material properties, withstand. Drain failure could result in significant over-load of
serviceability requirements such as deflection/deformation and the structure.
crack control as well as selection of appropriate ground and
lining stiffness values are items which can often result in 5.4 Design codes and analysis
ambiguity in the definition of applied design loads and may
dramatically affect the resulting performance of the tunnel Unfortunately, there is no known definitive Code identified for
structure. the design of underground openings. Generally designers use
There must always be a clear understanding of how the traditional building codes or other codes geared for above-
design has been undertaken and the construction sequencing ground structures and temper the approach and application of
must follow the design or the design must be re-evaluated to these codes based on underground design experience. Typical
conform to the works in the field. All too often the designer has codes that are applied to underground structures are: ACI 318,
predicated the analysis and design on conditions which are ACI 224R-01 (2008), AASHTO (2002), BS8110 (1997),
unrelated to the actual works being performed in the field. BS5400 (1988), Eurocodes (EN1992 2004; EN1997 2004) and
Hence, the design is not finished until the tunnel has been other such related design documents.
constructed. In 2000, the British Tunnelling Society (BTS) and the
Institution of Civil Engineers (ICE) issued the final draft of their
5.3.3 The role of geometry “Specification for tunnelling” (published by Thomas Telford)
Having the “User” defined requirements identified, there are This document refers to the following Eurocodes:
always select issues which need to be accounted for in the • Eurocode 1 (Basis of design and actions on structures,
design which will adequately allow sufficient space to construct 1994/95)
the final tunnel structure following the excavation and support • Eurocode 2 (Design of concrete structures, 1992)
phase for SES/NATM type of tunnels. • Eurocode 7 (Geotechnical design, 1996).
There are several major design decisions which must be These codes, subsequently updated as EN1990, EN1992 and
identified and resolved prior to the excavation phases such as: EN1997, have been used in a number of projects in the UK and
• Pressure design requirements (if not environmentally throughout the world.
required)
• Allowance for outer lining installation and deformations 5.4.1 Design code protocol
• User requirements such as: It is essential to define accurately what codes will be applied,
• Ventilation requirements along with where and how these codes will be used to design
• Alignment constraints the underground structure. Often misuse or alteration of the
• Water tightness criteria codes and design requirements result in an unfavourable
• Emergency egress and niches/laybys outcome during the design process. Upon the commencement of
• Constructability requirements for excavation and lining each project, it is extremely important that the protocol for use
installation of codes is clearly established and defined.
It is extremely important that there is a complete and full It is the desire of the authors that one day a proper code of
understanding of how these issues relate to the determination of practice will be developed for use in the design of underground
the tunnel excavation opening size. Once a tunnel excavation is structures and tunnels. Having an established code of this type
underway, it is very difficult and costly to have to increase the will have a distinct advantage of resolving arguments between
size of the existing opening as a result of any design errors or Designers, Owners and Contractors as the established code of
omissions. This issue has become more important as more practice will provide a sound baseline for the design process.
underground works have been following the design-build This aspect becomes ever more important as useable above
project delivery system to accommodate new demanding project ground sites become unavailable, thus promoting the
schedules. development and use of underground space.
More than other design works, there are often temporary
5.3.4 The role of maintainability structures that are required for the construction of the permanent
Regular inspection and maintenance are required to ensure structure. In the case of tunnels, this usually exists as braced
structures are performing as expected in achieving their excavations or the primary or outer-lining of tunnels constructed
anticipated design life, and to ensure any signs of distress or using the SES/SEM or NATM approach. The design protocols
deterioration can be diagnosed and addressed in a timely for these structures are no less important than that of the
manner. permanent or inner-lining. The only real exceptions to normal
The designer must consider how the cut and cover tunnel design requirements is in relation to the potential loading
will be inspected and maintained over the lifetime of the conditions likely to be encountered over the “useful” life of the
structure. For a reinforced concrete box type permanent structure and the requisite durability of the structure. Often
structure this is relatively straightforward. However, for cut and requirements for displacement and deformation are not nearly as
cover construction with integral support of excavation the finish stringent as that for the permanent structure provided that the
of the diaphragm wall for highway tunnels in particular may not initial excavation geometry has allowed for these movements so
be desirable from an Owner’s or user’s perspective. as not to encroach into the required opening for the final
In such cases one solution has been to disguise the irregular structure or cause distress to third party utility and land owners.
diaphragm wall finish behind a secondary finish wall. However, safety to the workers and to third parties must be
Obviously the area behind the finish wall is difficult to inspect maintained at all times and therefore the design of these
and maintain. However, the creation of sufficient space for man temporary works should be subject to similar codes of practice
access between the diaphragm wall and finish wall purely for to ensure that the structure(s) will perform satisfactorily.
inspection purposes is expensive. Therefore the secondary wall
is typically placed to leave an air gap of only a few inches
2912 B. Simpson et al. / State of the art Report: Analysis and Design

5.4.2 Design standards and codes behaviours of the surrounding soils and the interaction of the
The design standard selected for any given application can soils with the structure. Each of the software models and design
either be (a) a material code such as BS8110 or ACI318 which approaches has its own benefits and limitations as outlined
are principally building related standards, or (b) it can be related below.
directly to the tunnel function – whether highway or railway For the structurally based software the following findings are
tunnel – in which case in the USA the design would be in typical:
accordance with the American Association of State Highway • Models are simple and quick to develop and amend.
and Transportation Officials (AASHTO) or the American • It is relatively simple to model and simulate large numbers
Railway Engineering and Maintenance-of-Way Association of individual loads and combinations of loads, using both
(AREMA) standards respectively. service and ultimate load factors.
Table 5.1 indicates the differences in load factors for • Different factors can be applied to earth and hydrostatic
controlling loads for several design standards, each of which has loads based on the likelihood of a given design load to be
been used in the design of cut and cover structures. By exceeded.
inspection, the use of one listed standard versus another can • Models must also account for the variations in magnitude
result in significant differences. For example, earth pressure between vertical and horizontal loads as well as accounting
factors, which constitute one of the largest loads applied to the for the loading to be favourable or unfavourable for
structure can vary between 1.4 and 1.7, a difference of structure performance.
approximately 20%. • The effects of soil-structure interaction are modelled by
Note in particular: springs which do not represent the ground well. This often
a. AASHTO uniquely allows a factor of 0.65 to be applied for results in the models overestimating the structure forces and
checking positive moments in frames. hence the reinforcement requirements
b BS8110 Part 2 allows 1.2 to be used if applied to the “worst • It is difficult to model adequately the complex sequence of
credible” earth and water pressures. construction associated with the installation of the support
c The values for earth and, especially, water pressure are open of excavation system. This typically requires the
to some interpretation. introduction of ‘fake’ forces to the models to reproduce the
wall movements from the prior construction stage. This is
Table 5.1. Ultimate limit state load factors particularly important in situations where the support of
Lateral excavation is incorporated into the permanent structure.
Dead Live Hydrostatic With the geotechnical software the converse is true. For
Design Code Earth
Load Load Load tunnel design, given the points identified above, FEM may
Pressure
AASHTO 1.3 2.17 1.69a 1.3 provide a reasonable base to develop and understand the loads
to be used in the design of the tunnel liner. On larger and more
ACI 318 1.4 1.7 1.7 1.4 complex projects with multiple load types, a compromise is to
AREMA 1.4 2.3 1.4 1.4 use both modelling techniques. The structure software can be
used to develop multiple load cases and combinations to
BS 5400 1.15 1.5 1.5 1.5 identify which cause the most adverse conditions for stress. The
BS 8110 1.4 1.6 1.4b 1.4b controlling load combinations from the structural software can
be replicated using the geotechnically based software, thereby
Eurocode 2 1.35 1.5 1.35c 1.35c
promoting maximum design and construction economy.

While the adoption of any of the above listed standards 5.4.4 Examples of 3D analysis
results in a functional design, in each case load combinations, Yeow and Prust (2005) present an example of use of 3D finite
load factors, and material strength reduction factors exhibit element analyses to study the construction of a complex tunnel
differences which translate into different structure sizes, junction and its effect on existing tunnels. Figure 5.2 shows a
different reinforcement quantities and hence different detail of the junction, which was embedded in a mesh of
construction costs. 100,000 finite elements. The construction required about 90
It is therefore of fundamental importance that the designer stages of excavation. The ground movement was explicitly
understands the cost implications for the adoption of any one computed from the 3D tunnelling model without having to
standard versus another to identify the design standard which resort to the introduction of volume loss through stress
best reflects the function, durability and strength requirements relaxation or imposed volume change. Computed deformation
of any particular project in the most economic fashion. were used for damage assessment. Ability to prepare data for
analysis of this type and to carry it out in practicable timescales
5.4.3 Design analysis methods is still developing very rapidly. However, as discussed by
The majority of the analyses undertaken for design of cut and Yeow and Prust, providing adequate models of the soil and
cover structures and mined tunnels are now computer based. structural materials remains a major challenge.
Analysis models typically comprise two dimensional models Further examples are presented by Simpson et al (2006).
which adequately represent the linear nature of the tunnels. Figure 5.3 shows a situation at Stratford in East London, where
Three dimensional analyses are typically not necessary but are the new Channel Tunnel rain Link tunnel was to pass beneath
useful for modelling specific locations where special conditions exiting metro tunnels of the Central Line. Since it is known that
may exist such as rapid changes in cross section, large openings prediction of settlement due to tunnelling is extremely difficult,
or appurtenances such as adits, alcoves, and equipment rooms. a special method was devised in which an empirically derived
A number of software packages are available using either settlement field was imposed on a 3D finite element mesh
structurally based software such as Structural Analysis and which included the exiting tunnels. The details of the method
Design (STAAD) as manufactured by Research Engineers Inc are presented by Yeow et al (2005). Figure 5.4 shows the
which provide for two-dimensional plane frame analysis, or computed deformation of the existing tunnel which had a bolted
geotechnically based finite element/finite difference software steel lining, from which bolts were to be removed at various
such as Plaxis, manufactured by Plaxis BV or Fast Lagrangian points. It was predicted that if the design ground loss occurred
Analysis of Continua (FLAC) as manufactured by Itasca. in the new tunnels the existing tunnel would not bend much but
Whereas the structural software is obviously focused on the would shear at sections where the bolts were removed. In the
structural elements of the design, the geotechnical software event, actual ground loss and distortion of the existing tunnels
places greater emphasis on the simulation of the properties and were relatively small.
B. Simpson et al. / State of the art Report: Analysis and Design 2913

5.4.5 Example – sprayed concrete lining in clayey silt take up load. A particularly important decision is the choice of
The example shown in Figure 5.5 illustrates some of the overburden pressure, that is, whether any allowance is made for
problems to be considered in analysis of a sprayed concrete the type of arching action suggested by the finite element
lining. The internal tunnel dimension is approximately 11m analysis. Duddeck and Erdmann (1982) recommended that for
wide by 10m high, with a cover to diameter ratio (C/D) less shallow tunnels, C/D<3, a model without reduction of ground
than 2.0 It was excavated using an upper heading with pressure at crown is appropriate. In general, for shallow tunnels
temporary invert, with the bench excavated some distance (C/D < 2) the full overburden assumption is usually made.
behind the top heading. The primary lining of sprayed concrete Such an approach leads makes much greater demands on the
is 350mm thick, and the secondary lining of cast-in-situ capacity of the tunnel lining than does reliance on the results of
concrete is also 350mm thick. a finite element analysis of the type shown in Figure 5.8.
Figures 5.6 and 5.7 show two forms of analysis which have
been considered: finite elements and beam-spring models. The
2D finite element analysis uses a simple elastic-Mohr-Coulomb
model and takes the mesh through all the stages of tunnel
construction – heading and invert – with an appropriate
allowance for ground loss assumed to occur during the
excavation process. This inevitably leads to reduction in the
final stresses on the tunnel lining as the soil arches over the de-
stressed area of the tunnel, as illustrated in Figure 5.8. It is
assumed that in the long term the primary lining degrades and is
unable to carry bending stresses. If required, the formation of
plastic hinges in the linings, both primary and secondary, can be
modelled, though careful inspection of any plasticity which
implies cracking in the secondary lining is needed by designers
concerned about durability. Particular attention should be paid
to the rotation capacity at the cracked joint.

Figure 5.3: Intersection of the Channel Tunnel Rail Link and Central Line
tunnels at Stratford, East London.

Figure 5.4: Computed deformation of a Central Line tunnel.

Figure 5.5: Tunnel construction in clayey silt

Figure 5.2: 3D modelling of SCL tunnelling operation (after Yeow and


Prust 2005)

The spring model shown in Figure 5.7 can be used for an


analysis following the general procedures recommended by
Schulze and Duddeck (1964). This requires the designer to take
decisions about the stress field within which the tunnel is
placed, both in terms of the vertical and horizontal stresses,
though these are modified as the lining flexes and the springs Figure 5.6: Detail of 2D mesh at the tunnel
2914 B. Simpson et al. / State of the art Report: Analysis and Design

practices from the power and steelmaking industries


respectively.
The use of these replacement materials in reinforced
concrete has been demonstrated to offer a number of advantages
over Portland cement in terms of the durability and long term
performance of the concrete. These advantages include:
• a denser concrete mix which improves watertightness
• reduced requirement for mix water
• increased resistance to chemical attack
• reduced heat of hydration during setting and curing
The final bullet point is particularly significant. Lowering
the heat of hydration minimizes the temperature difference
between ambient air temperature and the peak temperature
within the concrete matrix during the setting process. By
minimizing the temperature differential the incidence of
problematic thermal cracking can also be minimized. This is of
Figure 5.7: Beam and spring model
particular importance for massive concrete structures such as
cut and cover structures which feature large concrete pours and
onerous conditions of restraint at wall/slab interfaces as thermal
cracks in theory will extend through the entire thickness of the
concrete section.
However, there are some disadvantages to the use of cement
replacement materials. The rate of gain of strength is slower
than for Portland cement concrete, which results in forms being
left in place for longer; material properties can vary
necessitating that all supplies come from a uniform source
where the properties are well understood. Analysis by Yazdchi
et al (2005) suggested that the speed of construction of a small
Figure 5.8: Principal stress plot showing arching
diameter tunnel with a sprayed concrete lining could be
restricted by lower rate of gain of strength.
A further issue for both types of analysis is to decide how
It can be seen from Table 5.2 that while there is extensive
factors of safety, particularly load factors, should be applied.
use of PFA in Hong Kong, Japan and the UK, in other major
As noted in Table 5.1, the codes require that different factors
construction markets considerable room for growth in the use of
are applied to dead and live loads. Some of them also
cement replacement materials exists.
differentiate between “unfavourable” and “favourable” dead
loads, applying a factor of 1.0 to favourable dead loads. These Table 5.2: PFA Production and Utilization 1995, after Meyer (2005)
distinctions are easily made if the factors are applied to the except UK data which is more recent, after Barnes and Sear (2004).
input data of the computations, but more difficult if an
unfactored analysis is carried out, with the intention of factoring Country Million Tons Million Tons % Utilization
the resulting bending moments and thrusts. Produced Utilized
Applying global factors to the forces derived from an China 91.1 13.8 15.1
unfactored analysis may significantly underestimate the adverse Denmark 1.3 0.4 30.8
effects of live loading upon the structure. For instance the
Hong Kong 0.63 0.59 93.7
transient effects associated with heavy rain upon partially
saturated soils may be underestimated if factors are simply India 57.0 2.0 3.5
applied to the loads derived from a calculation based on the soil Japan 4.7 2.8 59.6
properties in the partially saturated state. In that instance the
adverse combination of both a reduction in strength and Russia 62.0 4.3 6.9
stiffness of the soil and a temporary increase in load may not be USA 60.0 8.1 13.5
fully accounted for.
Fortunately, live loading is often less important to the design UK 5.7 3.1 55
of tunnel linings, but the distinction between favourable and
unfavourable dead loads can be controversial. Much of the In terms of the long term performance of underground
loading on a tunnel lining is due to the difference between structures the use of cement replacement materials is
vertical and horizontal stresses in the ground. If one of these is recommended. In addition, as natural resources become scarcer
deemed to be favourable whilst the other is unfavourable, the and more expensive, the expanded use of cement replacement
computed bending moments in the lining are increased materials should be promoted.
considerably. More common practice is to regard the ground The use of fibres, such as steel or polypropylene, in concrete
loading as coming from a “single source” (in the terms of for crack control and handling or to prevent explosive spalling
EN1990), so that the same load factor is applied to all earth of concrete under fire or an explosive event, respectively, is
pressures. If this is the case, and if live loading is unimportant, increasing throughout the world. The past performance in early
the effect of applying load factors to input data of the analysis, projects has supported the industry-wide use of these fibres and
or to the resulting bending moments, may be the same. as such, use of fibres is now given significant recognition and is
widely employed on modern day projects.
5.5 Materials and sustainability
5.5.1 Instrumentation and communications
In fairly recent years the use of cement replacement materials While there has been widespread use of geotechnical
such as ground pulverised fuel ash (PFA) and granulated blast instrumentation to monitor ground movements before and
furnace slag (GGBFS) has become commonplace in the during construction, with the advent of the internet has “real
specification of reinforced concrete for cut and cover and other time monitoring” – web based data provision and retrieval –
forms of tunnel. These materials are by-products from industrial many new systems are being developed everyday.
B. Simpson et al. / State of the art Report: Analysis and Design 2915

These systems offer a distinct advantage to construction of face of about 40m to 150m, a bench is taken out below the top
underground works in that immediate notification of adverse heading and supported. The final excavation of the floor, or
performance may be transmitted to the design/construction team invert, is taken out and supported, fairly close to the bench face,
24 hours a day. This level of monitoring and notification should depending on the ground conditions. The closed support ring
allow corrective actions to be implemented in a timely manner, around the full tunnel section comprises the outer lining, which
thus averting undesirable events. must stabilize the excavation before the inner lining is placed.
The outer linings generally have not been considered as part of
5.5.2 Performance monitoring the long-term or permanent support of the tunnels.
Tunnel excavation and lining installation should always include During excavation, survey displacement monitoring stations
both outer lining and inner lining monitoring programs. For cut- should be installed close to the excavation face, typically at
and-cover structures this would correlate to support of intervals of 10m to 20m, to measure the settlement and
excavation systems and the permanent structure, whereas convergence of the outer lining and provide verification of the
monitoring devices are located for not only the temporary works design of the outer lining. In poor ground conditions the spacing
but for the permanent works as well. By default, ground loading may be reduced to as little as 5m. Generally, each station
on the tunnel is directly related to the deformation of the ground consists of five targets, three in the top heading and two in the
and accurate and precise monitoring is one of the few tools bench. The monitoring results should be used to verify the
available to verify the design assumptions regarding load and design assumptions for the outer and inner linings. An
stability of the excavation. important aspect is that the information will be used to verify
The monitoring program can also include instrumentation to the assumed design ground loadings for the inner linings. Refer
better understand ground movement beyond the limit of the to Figure 5.9 for typical instrumentation layout.
excavation as well as loading on the outer and inner liner.
Distinct instrumentation stations that are representative of
sections of expected ground types (classes) are extremely
helpful in verification of anticipated tunnel performance as well
as for areas that have been determined to be difficult or
problematic.
It is advantageous that the Designer specifically requires that
these instrumentation stations are installed and that locations of
installation are confirmed and agreed to by the Designer.
Regular reading, interpretation and evaluation of the stations is
also paramount. In mined tunnels, often the full ground load
resulting from the excavation is not evident until the drive face
is 4-6 diameters past any given location.
Predicted deformations of the outer lining as provided by
analysis may be selected and used as “control” or “target”
values during the excavation whereby these are compared to the
measured deformations from the monitoring stations.
Depending on the ratio of the calculated deformations to the
measured ones, certain pre-agreed actions are implemented
when the ratio reaches certain values. For example, “alert level
1” occurs if the ratio reaches 1/3, “alert level 2” if 2/3, and
“alert level 3” if 1.0. The control values are determined for each
stage of the excavation thus allowing for deformations to occur
for each of the 3 standard stages of excavation, heading, bench
and invert. A typical relationship between the “Control Criteria” Figure 5.9 – Instrumentation Station
and “Safety Control System” is provided in Table 5.3.
5.5.3 Constructability reviews
Constructability reviews are as essential to the design as are the
Table 5.3 Lining Performance Control Criteria drawings, specifications and supporting calculations. All the
drawings and calculations will be for nothing if the design is not
CONTROL
ALERT LEVEL
VALUE
ACTION constructible. As an example, if difficult ground is encountered
the tunnel invert excavations should be made deeper and more
Standard Operations - Standard Measurements circular; this should be anticipated in cases where ground
conditions are not well known or if it is known that poor ground
Increase Measurement conditions are likely to exist. This sort of flexibility in the
1/3 Control Frequency, Conduct Site design should be considered in the beginning and properly
Level I Alert
Value Inspections, Issue Strict planned for, including layout of plant and equipment. The
Work Instructions general concept of excavation equipment should be flexible to
Strengthen Measurement meet the changing demands of underground work during the
2/3 Control System, Carry Out / Perform tunnel construction.
Level II Alert
Value Minor Control Works Optimized reinforcement, especially in the arch lining,
(“Auxiliary Measures”) benefits the inner lining. High reinforcement content will
Stop Excavation, Analyze increase the probability of honeycombing and voids. These
Control Cause and Tendency of reviews are often performed prior to tender award but they also
Level III Alert
Value Deformation, Select Tunnel continue through the construction phases as well.
Reinforcement Measures The beginning of mined tunnelling should always be based
on a conservative design. The observations during early site
In view of the large excavation sizes, the tunnel cross-section execution should accelerate the learning curve and optimize
is usually excavated in stages. Initially a top heading with an support requirements. As information is gathered, the Tunnel
invert on or just above the tunnel spring line is excavated and Construction Engineer and the Lead Geotechnical Engineer can
stabilized, if necessary with a temporary invert of shotcrete to
form a closed support ring. At a variable distance behind the
2916 B. Simpson et al. / State of the art Report: Analysis and Design

work closely with the Designer who can then refine the analysis or curing on an annual basis (Transportation Research Board
to support the decisions for lining design and support selection. 2007, Mehta 2004, Meyers 2005).
Precise Monitoring of deformation and interpretation of the In addition to the depletion of these natural resources, the
deformation will help identify the best support solutions to production of cement in itself expends considerable amounts of
minimize risk of collapse. However, deformations alone do not fossil fuel and electrical energy. Annual production of cement is
always tell the full story. Full instrumentation stations which are approximately 1.6 billion tons, and is expected to exceed 2
capable of identifying ground mobilization and lining stress billion tons by 2010. The energy expended on the creation of
allow for a full understanding of the interaction between ground one ton of cement generates an equivalent weight of carbon
strains and lining performance. dioxide (CO2), a principal contributor to global warming. The
annual production of 1.6 billion tons (and rising) of CO2
5.6 Future developments corresponds to approximately 7% of the global emission of this
gas.
In this emerging era of climate change and constrained In addition, concrete construction debris from demolition
resources it is incumbent on the engineering profession to constitutes a large percentage of solid waste disposal. In North
design for value through economic use of materials and the America, Europe and Japan concrete and masonry rubble
promotion of engineered solutions which are durable and accounts for approximately two thirds of all construction and
sustainable. demolition waste. It is reported that in excess of 1 billion tons of
construction and demolition waste is generated each year
5.6.1 Design standards and codes (Mehta 2001).
Development of a standard specific to below grade construction Clearly none of these practices is sustainable in the long
which is consistent with construction practices (or a number of term. However, solutions to mitigate or at least reduce their
standards dependent upon regions/jurisdictions) will help the impacts are being identified and implemented. These solutions
review and acceptance of designs and construction performance. include the increased use of cement replacement materials – as
As an example, appropriate load factors for groundwater identified earlier, these materials are significantly underutilized,
levels are an important element of the design when hydrostatic and the use of recycled concrete aggregates should be promoted.
pressures form a large part of the permanent load. The larger the Typically concrete mixes are specified to contain roughly
percentage of total load resulting form water pressure should 30% of cement replacement material. With the benefits
merit high scrutiny of the selection of load factor to be used described previously proponents of the use of pfa/ggbfs in
with that load. Long term creep and concrete strain need to be concrete production advocate that this percentage should be
reviewed in relation to the load applied.. significantly higher and a content of 50-60% cement
Accordingly the UK National Annex to Eurocode 7 replacement should be sought. This increased volume of cement
allows/encourages “direct assessment” of design values of replacement materials will reduce requirements for cement
ground water pressure (implying that no further factor is to be production and correspondingly reduce CO2 emissions.
applied to these): An additional benefit of increased usage of cement
The partial factors specified in the National Annex to BS EN replacement materials is the reduced requirement for mixing
1990:2002 [ie the normal partial factors on loads] might not be water. Concrete mixes using high percentages of cement
appropriate for self-weight of water, ground-water pressure and replacement material have been demonstrated by testing to
other actions dependent on the level of water, see 2.4.7.3.2(2). require 20% less mix water than corresponding mixes which are
The design value of such actions may be directly assessed in purely cement based. In many regions of the world water is a
accordance with 2.4.6.1(2)P and 2.4.6.1(6)P of BS EN 1997- precious resource. A 20% reduction in concrete mix-water
1:2004. Alternatively, a safety margin may be applied to the would equate to approximately 200 billion gallons of water
characteristic water level, see 2.4.6.1(8) of BS EN 1997-1:2004. saved per year. This is undoubtedly a significant volume (Mehta
An alternative approach is to calculate structural stresses for 2004).
unfactored “characteristic” water pressures and apply factors to
the structural stresses. There is considerable merit in checking 5.6.4 Contracting practices
both approaches in a design. Due to the complexities involved with underground
Accordingly design standardization and consistency of construction in an urban environment, it is time to move away
application of load factors/combinations etc. are very important from low bid construction which is prevalent over much of the
aspects of design for underground structures and further world to a ‘best value’ approach whereby the contractors
development of the codes for design of such structures is technical proposal forms an official part of the bid evaluation
warranted. process (Pollalis 2006).
Consideration should be given to early contractor
5.6.2 Analysis involvement in the design process, using a negotiated or target
More widespread integration of the temporary support of price contract approach as was successfully implemented on the
excavation (SOE) into the permanent structure helps to reduce UK Channel Tunnel Rail Link (Cathart 2005).
environmental impacts associated with the work as it lessens the Under such a scenario qualified contractors would bid on
tunnel footprint and minimizes impacts upon utilities, and design documents complete to a preliminary engineering level
expensive real estate acquisition. of detail. The selected contractor would subsequently work with
Soil-structure interaction models can be used to demonstrate the owner and designer to complete the design engineering.
the ability of SOE in the long term condition to relieve load on Through this process, the designer understands and can design
the cast in place structure, and correspondingly reduce structure to accommodate the contractor’s proposed means and methods,
size and cost. Use of integral SOE should be studied further to the contractor provides constructability review of the design,
assist in the use of a hybrid solution for the permanent structure, and the owner receives a bid price with reduced contingency.
while avoiding potential durability issues. Once the design is finalized, the owner and contractor
renegotiate the contract price based upon the improved
5.6.3 Materials understanding and development of the project scope. In the
It is well documented that the concrete industry is one of the event that an agreement cannot be reached, the parties can
largest consumers of natural resources – relatively recent terminate negotiations, the contractor is paid for his design
statistics suggest that the concrete industry consumes over 10 effort, and the contract can be rebid.
billion tons of sand and aggregates and 1 billion tons (1 trillion
gallons) of water, not including water for wash down of mixers
B. Simpson et al. / State of the art Report: Analysis and Design 2917

6 SEISMIC THEMES OF GEOTECHNICAL expanded. Two such topics are highlighted in this paper. The
ENGINEERING INTEREST: BEYOND THE STATE OF emphasis is on elucidating the analysis of seismic soil-
PRACTICE foundation interaction in the presence of large soil deformations
and near-failure conditions.
It is emphasized, however, that many of the ideas outlined in
6.1 Introduction: topics that have emerged from recent the following sections, although supported by some field
earthquakes evidence, are not yet to a sufficient extent developed to be
directly applicable in design. Many more analytical studies and
This part of the report addresses issues related to what is called several large-scale or centrifuge experimental tests, as well as
“performance-based design” of foundations against two additional well-documented field evidence, are needed before
earthquake−related hazards: (i) emergence of the rupture of a these ideas are turned into reliable design methods accepted by
seismic fault underneath a structure, and (ii) strong dynamic the profession.
shaking resulting from seismic waves emanating from the whole
rupturing fault. Against the latter hazard (i.e. strong ground 6.2 Foundations interacting with a rupturing seismic fault
shaking) emphasis will be on the mobilisation of bearing
capacity mechanisms for slender structures on shallow
foundations. The topics chosen in this chapter have been 6.2.1 Problem and motivation
prompted by observations in many recent earthquakes, most “Strong” structures founded on the surface of, or at depth in,
importantly from Northridge (1994), Kobe (1995), Kocaeli soil have often resisted successfully the loading induced by a
(1999), and Chi-Chi (1999). Valuable lessons of great rupturing seismic fault (Duncan & Lefebvre, 1973; Bray 1990).
significance have emerged from these events. Here are a few of In the three Turkey and Taiwan earthquakes of 1999 numerous
them: structures (buildings, bridge piers, retaining structures,
• Structures on conventionally stiff raft and box foundations electricity pylons, dams, tunnels) were located directly above
can survive dip-slip and strike-slip fault ruptures producing the propagation path of the rupturing faults ⎯ strike slip,
offsets of the order of 1m beneath the structure. By normal, reverse. Some of these structures performed
contrast, pile foundations are directly affected by the rupture remarkably well; others failed dramatically. These observations
and, moreover, tend to impose on the super-structure the had a strong motivating influence to modify the pertinent
fault-induced differential displacements; both structural and clauses of seismic codes and to conduct further research. For it
pile damage are thus likely to occur. became immediately clear that the strict prohibition of building
• Even on very soft soil, a slender structure subjected to in the immediate vicinity of active faults, which the prevailing
strong shaking may undergo severe rocking oscillations, seismic codes have demanded, was unduly restrictive if not
meaningless.
accompanied by uplifting. Large deformations are thus
In addition to several geologic factors that contribute to such
induced in the soil and bearing capacity mechanisms are
behaviour, the role of a soil deposit that happens to cover the
mobilised alternately under each half of the foundation. But
rock appears to be significant. If, where, and how large will the
the structure does not necessarily topple, and perhaps it does dislocation emerge on the ground surface (i.e. the fault will
not even suffer detrimentally-large residual rotation and outcrop) depends not only on the style and magnitude of the
displacement. fault rupture, but also on the geometric and material
• Geotechnical systems yielding only in one direction, such as characteristics of the overlying soil deposits. Field observations
gravity retaining walls and slopes, are quite sensitive to the and analytical and experimental research findings (Bray et al,
nature and direction of seismic shaking and are thus greatly 1994a, b; Cole & Lade, 1984; Lade et al, 1984; Anastasopoulos
affected by phenomena such as forward-directivity and et al 2007) show that deep and loose soil deposits may even
fling-step, often observed close to the seismogenic fault. mask a small-size fault rupture which occurs as a dislocation at
• Conventionally designed caisson quay-walls prove to be their base; whereas, on the contrary, with a cohesive deposit
rather robust structures against toppling but they may especially of small thickness, a large offset in the base rock is
experience (large) seaward displacement and rotation if the likely to cause a distinct fault scarp of nearly the same
foundation soil is very deformable or, even worse, displacement magnitude. One important finding of the above
liquefiable; by contrast, the retained soil does not liquefy studies is that the rupture path in the soil is not a simple
(contrary to a widely-held misconception) and will only extension of the plane of the fault in the base rock: phenomena
exert moderate forces on the wall, of the same order as the such as “diffraction” and “bifurcation” affect the direction of
pseudo-statically determined Mononobe-Okabe active the rupture path, and make its outcropping location, the offset
forces. magnitude, and the shape of the deformed surface difficult to
• End-bearing pile foundations can perform very well in predict.
“level” ground, even if the surrounding soil were to liquefy
massively, as long as there is no danger of buckling (usual
case with the present-day large diameters of cast-in-place
piles). However, in “non-level” ground, liquefaction
triggers soil flow; as a consequence, piles could undergo
severe “kinematic” lateral deformations and possibly
bending failure. (“Kinematic” refers to ground-
displacement induced stressing of the pile along its depth, as
opposed to “inertial” loading which originates at the
superstructure and is then transmitted on the top of the
piles.) The possibility of buckling of small diameter end- Figure 6.1: Configuration of the soil–foundation system subjected to a
bearing piles cannot be dismissed, however, as has been normal fault dislocation at the base rock.
pointed out by Bhattacharya et al (2004, 2005) and Kerciku
et al (2007). Our main interest here is to study how a structure sitting on
Many theoretical and experimental studies have been top of the fault breakout behaves. It turns out that an interplay
initiated following these field observations, so that today the occurs between the propagating fault rupture, the deforming
range of topics of interest in Seismic Geotechnics has greatly foundation soil, the differentially displacing foundation, and the
2918 B. Simpson et al. / State of the art Report: Analysis and Design

supported structure. Two different phenomena take place.


First, the presence of the structure modifies the rupture path.
Depending on the rigidity of the foundation and the weight of
the structure, even complete diversion of the fault path before it
outcrops may take place. Obviously, the damage to a given
structure depends not only on its location with respect to the
fault outcrop in the “free-field”, but also on whether and by how
much such a diversion may occur. Second, the loads
transmitted from the foundation on to the soil tend to compress
the “asperities” and smoothen the “anomalies” of the ground
surface that are produced around the fault breakout in the
free−field, i.e. when the structure is not present. Thus,
depending on the relative rigidity (bending and axial) of the
foundation with respect to the soil, as well as on how large the
structural load is, the foundation and the structure experience
differential displacements and rotation, different from those of
the free-field ground surface. This phenomenon, given the Figure 6.2: Finite element discretisation and the two steps of the
name Fault Rupture−Soil−Foundation−Structure Interaction analysis: (a) fault rupture propagation in the free-field, and (b) interplay
between the outcropping fault rupture and the structure (termed Fault
(FR−SFSI) is briefly elucidated below. Rupture–Soil–Foundation–Structure Interaction, FR-SFSI).
6.2.2 Shallow foundations on top of rupturing fault A typical result showing the interplay between loose (ID =
The problem studied here is illustrated in Figure 6.1. A uniform 45%) soil, rupture path, and a perfectly rigid foundation
soil deposit of thickness H at the base of which a normal fault, carrying a 3-storey structure is given in Figure 6.2. A base rock
dipping at an angle Į (measured from the horizontal), produces dislocation of 2 m (5% of the soil thickness) is statically
downward displacement (called “dislocation” or “offset”) of imposed. The structure is placed symmetrically, straddling the
vertical amplitude h. Note that the movement of the fault free-field fault breakout (i.e. the foundation is placed with its
during an earthquake is in one direction only (and not middle coinciding with the location where the fault outcrops in
oscillatory) and takes place rather slowly (on the order of tens the free field). A distinct rupture path (with high concentration
of seconds rather than one-tenth of a second) − Ambraseys and of plastic shearing deformation and a resulting conspicuous
Jackson (1984). The analysis is static and is conducted in two surface scarp) is observed only in the freefield. The presence of
steps. First, fault rupture propagation through soil is analysed the structure with its rigid foundation causes the rupture path to
in the free field, ignoring the presence of the structure. Then, a bifurcate at about the middle of the soil layer. The resulting two
strip foundation of width B carrying a particular superstructure branches outcrop outside the left and near the right corner of the
is placed on top of the free−field fault outcrop at a specified foundation, respectively. The soil deformations around these
distance S (measured from its corner), and the analysis of branches are diffuse, and the respective surface scarps are much
deformation of the soil–structure system due to the same base milder than in the free field. Thanks to the substantial
dislocation h is performed. The analyses are conducted under weight of the structure and the flexibility of the ground, the
2-D plane-strain conditions ⎯ evidently a simplification, in structure settles and rotates as a rigid body. The foundation
view of the finite dimensions of a real structure in the direction does not experience any loss of contact with the ground;
parallel to the fault. A limited number of 3-D analyses have apparently, the foundation pressure is large enough to eliminate
shown that the 2-D approximation leads to slightly conservative any likely asperities of the ground surface. As a result of such
results (i.e., it leads to somewhat larger rotation) (Gazetas et al behaviour, the structure and its foundation do not experience
2007). any substantial distress, while their rotation and settlement
Among several alternatives that were explored, the FE model could be acceptable.
shown in Figure 6.2 produced results in excellent accord with Several tectonic, geometric, and material factors affect the
several centrifugal experiments conducted at the University of interplay between an emerging fault rupture, the soil, and the
Dundee for both steps of the analysis (Anastasopoulos et al foundation. For a fairly detailed parameter investigation
2007, 2009). A parametric investigation revealed the need for a relating to the subject, see Anastasopoulos et al (2007, 2009).
long (B ≈ 4H) and very refined mesh (element size of 0.5 m − This presentation will consider the consequences of this
1.0 m) along with a suitable slip-line tracing algorithm in the interplay for a rigid mat foundation, 20m wide, carrying a 2-
region of soil rupture and foundation loading. An elastoplastic story frame structure, and resting on top of a 40m thick deposit
constitutive model with the Mohr-Coulomb failure criterion and of dry sand. Two different densities are considered
isotropic strain softening was adopted and encoded in the parametrically: ID = 45% (loose sand) and ID = 80% (dense
ABAQUS finite element environment. Similar models have sand). Typical values of (peak and critical-state) angles and
been successfully employed in modelling the failure of dilatancy angle are assigned to each density, in addition to some
embankments and cut slopes (Potts et al, 1990). Modelling other secondary modelling differences. For ID = 80%: ijp =
strain softening was shown to be necessary; it was introduced 45Ƞ , ijcs = 30Ƞ , ȥp = 18Ƞ, ȥcs = 0o. For ID = 45%: ijp = 32Ƞ ,
by suitably reducing the mobilised friction angle ijmob and the ijcs = 30Ƞ , ȥp = 3Ƞ, ȥcs = 0o. Regarding their stiffness, both
mobilised dilation angle ȥmob with increasing plastic octahedral soil deposits are taken as "Gibson” type soils, i.e. with elastic
shear strain. With all the above features, the FE formulation is Young’s modulus increasing linearly with depth: E = 5z (MPa,
capable of predicting realistically the effect of large m) for ID = 80% and E = 2z (MPa, m) for ID = 45%. The style
deformations with the creation and propagation of shear bands. and magnitude of the fault rupture play an important role in the
The foundation, modelled with linear elastic beam elements, response (and eventually the survival) of the structure. A
is positioned on top of the soil model and connected to it normal faulting is considered here, emerging with an angle Į =
through special contact elements. The latter are rigid in 45o on the surface of the base rock and with a dislocation
compression but tensionless, allowing detachment of the (offset) having a vertical component h = 2m, that is 5% of the
foundation from the bearing soil (i.e. gap formation beneath the overlying soil thickness (H =40m).
foundation). The interface shear properties follow Coulomb’s Another important factor is the exact position of the
friction law, allowing for slippage. Both detachment and foundation with respect to the outcropping location of the fault
slippage are important phenomena for realistic foundation rupture on the ground surface. Since this location is affected by
modelling. the foundation-structure itself, we use as reference the point O
B. Simpson et al. / State of the art Report: Analysis and Design 2919

in Figure 6.1, where the fault outcrops at the free field ground bending moment ⎯ this indeed constitutes the major
surface. Having numerically determined point O, we specify as distress of the foundation from the rupturing fault. The
the position of the foundation the distance S of its left edge (i.e. survival of the structure depends on its ability to safely
the edge on the hanging wall) from O. sustain this moment.
Two values of S are considered here: S = 4m and S = 16m. 3. In dense soil, the observed asymmetric reduction in the
In the former case, the fault, if unaffected by the foundation, contact area leads to an increase of the average normal
would have emerged near the left edge of the foundation; 80% contact pressure (to about 1.25 po) and to an unavoidable, if
of the foundation would have been located on the stable small, rotation of the foundation. All this culminates in a
footwall. In the latter case, the fault would have emerged near nearly triangular distribution of normal contact pressures
the right edge of the foundation; only 20% of the foundation with a peak of about 3po near the edge of the fault scarp,
would have been located on the stable footwall. Figures 6.3 and while the right edge of the foundation starts to uplift from
6.4 illustrate the response of the foundation in these two cases, the soil. This generates an additional cantilever at the right
for each of the two soil densities. (In both figures the vertical edge and further aggravates the rigid-body rotation of the
scale is substantially exaggerated; the slopes thus appear much foundation. This constitutes the second, “operational”
steeper than in reality.) It is noted that in all four cases distress of the foundation from the rupturing fault. Note,
examined the foundation is lightly loaded (mean contact however, that for large enough values of po the local soil
pressure po = 30kPa). The influence of po on the behaviour of yielding will reduce the peak and lead to a more uniform
the foundation-structure system will be explored later. pressure distribution.
Several trends are worthy of note in the two figures: 4. On loose soil the size of the gap is restricted to 2.5 m, while
1. For the first case, of the fault emerging near the left edge (S along the remaining 17.5 m of the foundation width full
= 4m, or S/B = 0.20), this lightly loaded foundation causes contact is maintained. Hence, the bending moment of the
only a minor diversion of the rupture path from its free-field cantilevered part is only 40% of the bending moment that
position (Figure 6.3). This diversion is clearly noticeable develops under the same conditions on the dense soil. This
only in loose soil (deviation to the left of about 1.5 m away favourable behaviour stems apparently from the greater
from the footing centre, i.e. towards the hanging wall). compressibility of the loose soil and the ensuing depressing
2. A profound consequence of the emergence of the fault of the fault scarp, as is evident in the figure. However, an
under the foundation is the development of gaps between unfavourable consequence of the increased soil
soil and foundation. On the dense soil nearly the whole left compressibility is the greater (by a factor of nearly 3)
part of the foundation ( ≈ 4 m long) turns into a cantilever. rotation of the rigid foundation.
The structural loads above this part induce a (substantial)

Figure 6.3: Response of soil surface and foundation due to a normal fault emerging at a distance S = 0.20B from the left
edge under the foundation. (Vertical scale exaggerated.)

Figure 6.4: Response of soil surface and foundation due to a normal fault emerging at a distance S = 0.80B from the left
edge under the foundation. (Vertical scale exaggerated.)
2920 B. Simpson et al. / State of the art Report: Analysis and Design

2. For the second case (Figure 6.4), of the fault emerging soil. Such a good response of a building founded on loose soil
near the right edge (S = 16m or S/B =0.80), a substantial on the hanging wall is reminiscent of several success stories
diversion of the fault takes place in both loose and dense from the Kocaeli 1999 earthquake, especially of the building
sand: the fault deviates by 4 m towards the footwall in in Denizevler across the entrance from the Ford factory, near
both cases, emerging beyond the foundation at the right- Gölcük (see Anastasopoulos et Gazetas 2007a). Figure 6.5
hand side. This significant theoretical finding has been shows in some detail this building and the geometric
verified with several observations of actual behaviour in characteristics of fault outcropping.
the Kocaeli 1999 earthquake (Anastasopoulos & Gazetas The above results are all for a specific small contact
2007a,b). But the similarity in behaviour between loose pressure, po = 30 kPa, at the foundation−soil interface, typical
and dense sand ends here. Significant differences are of a 2 story building. Although not shown here, the effect of
noted between the two cases: an increase in the number of stories is quite beneficial on
• The fault scarp that is formed near the right edge of the loose sand, and somewhat less on dense sand. The most
building is far more conspicuous in loose sand. significant benefits are the decrease of foundation rotation
• In dense sand the fault rupture undergoes bifurcation, (and thereby of building tilting) and the elimination of a large
with the secondary rupture branching to the left and part of uplifting. As a consequence, the survival of a heavy
emerging underneath the structure, not far from its centre. building founded on loose soil above a major fault rupture
• As a result, on dense sand, the middle part of the building seems possible, in qualitative accord with numerous such
loses contact with the bearing soil for about 10m, while success stories in several earthquakes.
the left and right part remain in contact for about 2m and The significance of the magnitude of the (average) contact
pressure, po, is summarized in Figure 6.6. The figure gives
7m, respectively.
parametric results for the size of the unsupported (detached)
On loose sand, the response of the foundation is quite
regions, u, and the effective (in-full-contact) regions, b, of the
favourable: not only is the dislocation diverted and outcrops
foundation. Both are normalized by the width B of the
beyond the right edge as already noted, but full contact is
foundation. In addition to po, the examined variables include
maintained over the whole length of the soil−foundation
the location of the foundation with respect to the free-field
interface. The distress of the foundation is thus significantly
fault outcrops, S, and the relative density of the soil, ID.
less on loose than on dense sand. Also smaller on loose sand
is the (rigid-body) rotation of the foundation, thanks to the
larger depressing of the fault scrap on the more compressible

Figure 6.5: The remarkable performance of a 5-story building near Gölcük, Turkey, “on top” of the normal fault rupture in the Kocaeli 1999
Earthquake: (a) photograph showing the fault being directed towards the building; (b) photograph showing the vertical displacement
reaching 2.3 m, along with a horizontal component of 1.1 m, measured on the torn-apart fence of the building; (c) simplified cross-section of
the building and the fault; and (d) plan view of the foundation (box-type foundation with cross tie beams), along with the horizontal
displacements measured around the building, and a sketch of the diverted fault. (after Anastasopoulos & Gazetas, 2007a).
B. Simpson et al. / State of the art Report: Analysis and Design 2921

Figure 6.6: Summary of parametric results showing in normalised form which parts of the foundation remain in full contact
(constituting the effective width, b, or b1 + b2 in certain cases) or separate from the soil (constituting the unsupported width, u,
or u1 + u2 in certain cases). The fault ruptures at three different locations: (a) s = 0.2B, (b) s = 0.5B, and (c) s = 0.8B. The effect
of surcharge load, po, is shown for the dense (left) and the loose (right) soil deposit. Indicative cross-sections of the foundation
on dense and loose sand are shown on the left and on the right, respectively, showing the no-contact areas and crude sketches of
the soil reactions.
unavoidable uncertainty on the exact location of the emergence
The following trends are noted: of any fault, this means avoiding to build “in close proximity”
• On dense sand, for S/B = 0.2 the increase of po leads to an to the fault
increase of the effective foundation width b from 0.60B to
0.75B, reducing the maximum (unsupported) cantilevered 6.2.3 Pile and Caisson Foundations
spans from u1 = 0.25B to 0.20B (on the left side) and from The role of piles in supporting structures straddling seismic
u2 = 0.2B to 0.10B (on the right side). For S/B = 0.5 there faults is far from clear. Circumstantial evidence from recent
develop two unsupported spans: one (u2) cantilevered on the earthquakes has implicated the piles in some structural damage
right side (towards the hanging wall) and the other (u1) — see for example the analysis of the damage of the pile-
doubly-supported under the left half of the foundation; thus supported Attaturk Stadium in Denizevler during the Kocaeli
there exist two areas of contact: a small (b1) and a larger Earthquake (Anastasopoulos & Gazetas 2007). Systems “tied”
(b2). The increase of po leads to increasing b2 and to the different blocks of the fault may indeed be vulnerable.
decreasing u1 and u2 , while b1 remains essentially the same. Reference is made to Gazetas et al (2007a,b) for a more detailed
exposition of the behaviour of piled foundations under a normal
In stark contrast, in the case of S/B = 0.8 the increase of po
fault rupture.
does not seem to play a significant role.
Rigid caisson foundations are clearly advantageous. The
• On loose sand, the effective width is invariably much larger faulting-induced deformation causes a more-or-less rigid-body
than on dense sand. The most noteworthy effect of displacement and rotation that are in general smaller than those
increasing po is in the case of S/B = 0.5: the unsupported of surface or piled foundations. Reference is made to
spans disappear and full contact is established for po > Anastasopoulos et al (2008a) for more information on the
60kPa. But the biggest beneficial effect of increasing po is subject and an application to an actual bridge problem in
the reduction of the rotation experienced by the foundation. Greece.
The engineer could use Fig. 6.6 to preliminarily design the
foundation raft. But this should be done only if he cannot
reliably avoid building “on the fault”. In view however of the
2922 B. Simpson et al. / State of the art Report: Analysis and Design

6.3 Slender structures on shallow foundations: mobilisation especially uplifting from, their base during oscillatory
of bearing capacity mechanisms seismic motion has been a key to their survival. These
phenomena cannot therefore be ignored.
• Compatibility with state of the art structural earthquake
6.3.1 Conventional wisdom and the need for change engineering is another reason to compute the complete
Seismic design of structures recognises that highly inelastic inelastic lateral load−displacement or load−rotation
material response is unavoidable under the strongest possible response of the foundation system, to progressively
shaking. Displacements as large as 3 times the yield increasing loads up to collapse. Otherwise the
displacement (in earthquake terminology “ductility” of 3) or “performance-based” structural analysis, will be incomplete.
more are usually allowed to develop under seismic loading. This
It is therefore logical to extend the inelastic analysis to the
implies that the strength of a number of critical bearing
supporting foundation−soil system.
elements is fully mobilised. In the prevailing structural
terminology “plastic hinging” is allowed as long as the overall
structural stability is maintained.
By contrast, a crucial goal of current practice in seismic
“foundation” design, particularly as entrenched in the respective
codes (e.g. EC8), is to avoid the mobilisation of “strength” in
the foundation. In structural terminology: no “plastic hinging”
is allowed in the foundation, below the ground surface.
Thus the conventional approach to seismic foundation design
introduces factors of safety against sliding and against
exceedance of ultimate capacity, in a way similar to the
traditional static design. This approach involves two
consecutive steps of structural and foundation analysis:
a. Dynamic analysis of the structure is first performed, in
which the soil is modelled as an elastic medium represented
by suitable translational and rotational springs (and,
sometimes, with the associated dashpots). The dynamic
forces and moments transmitted onto the foundation are
derived from the results of such analyses, after the column
forces have been reduced by dividing with a ductility-
capacity dependent factor. (In EC8 this factor is called
“behaviour” or “q” factor. It varies between 1 and 4;
depending on mainly the ability of the superstructure to
undergo safely large inelastic deformations.)
b. The foundations are then designed in such away that these Figure 6.7: Top: problem geometry. Bottom: the two studied
foundation solutions.
transmitted horizontal forces and overturning moments,
increased by “overstrength” factors, would not induce
sliding or bearing capacity failure. 6.3.2 Towards a new design philosophy: “plastic hinging” in
The use of “overstrength” factors is necessitated by the so- soil-foundation systems
called “capacity design” principle, under which plastic hinging Excluding structural yielding in the isolated footing or the
is allowed only in the structural elements — not in the foundation beam, three types of nonlinearity can take place and
below−ground (and hence uninspectable) foundation and soil. modify the overall structure–foundation response:
Therefore, structural yielding of the footing and mobilisation of a. Sliding at the soil−foundation interface. This would happen
bearing capacity mechanisms is not allowed. However, there is whenever the transmitted horizontal force exceeds the
a growing awareness in the profession of the need to consider frictional resistance. As pointed out by Newmark (1965),
soil-foundation inelasticity, in analysis and perhaps even in thanks to the oscillatory nature of earthquake shaking, only
design (see: Pecker (1998), Paolucci (1997), Martin & Lam short periods of exceedance usually exist in each direction;
(2000), Allotey & Naggar (2003)). This need has emerged hence, sliding is not associated with failure, but with
from: permanent irreversible deformations, as will be shown in the
subsequent section of this paper.
• The very large accelerations and velocities recorded in
b. Separation and uplifting of the foundation from the soil.
earthquakes, which would impose enormous inelastic
This happens when the overturning moment tends to
demands to structures if soil−foundation “yielding” would produce net tensile stresses at the edges of the foundation.
not effectively limit the induced accelerations. Thanks again to the oscillatory nature of the seismic
• Seismic retrofitting of a structure increases the shear shaking, the ensuing rocking oscillations in which uplifting
capacity of some elements and hence the forces onto their takes place do not lead to overturning of the structure.
foundations; it might not be feasible to undertake them There is not detriment to the vertical load carrying capacity
elastically. A (stiff) concrete shear wall inserted to upgrade and the consequences in terms of induced vertical
a frame carries most of the inertia-driven shear, and thereby settlements may be minor. Moreover, in many cases,
transmits a disproportionately large horizontal force and footing uplifting is beneficial for the response of the
overturning moment onto the foundation. If uplifting, superstructure, as it helps reduce the ductility demands on
sliding, and mobilisation of bearing capacity failure columns.
mechanisms are correctly taken into account, the shear wall c. Mobilisation of bearing capacity failure mechanisms in the
“sheds” off some of the load onto the columns; the opposite supporting soil. Such inelastic action is almost unavoidable
is erroneously the case when such inelastic action is with uplifting of the foundation. In static geotechnical
disallowed. analysis large factors of safety are introduced to ensure that
• Many slender historical monuments have apparently bearing capacity modes of failure are not even approached.
survived several strong seismic motions in their (often long) In conventional seismic analysis, bearing capacity is
life. While under static conditions they would have easily avoided thanks to an “overstrength” factor of about 1.40.
toppled or otherwise failed, it appears that sliding at, and (Note that this factor multiplies the (maximum) design
B. Simpson et al. / State of the art Report: Analysis and Design 2923

moment at the base of the superstructure; it is this increased 6.3.3 Results of a comparative study
moment that must be carried by the foundation–soil To illustrate the interaction between soil, foundation, and
system.) The oscillatory nature of seismic shaking, structure under strong seismic shaking mobilising inelastic
however, allows again the mobilisation (for a short period deformations in the soil we have selected the system
of time) of the maximum soil resistance along a portrayed in Figure 6.7. A single-column concrete bridge
continuous (“failure”) surface. No collapse or pier, 3m in diameter and about 12m high, carries a deck load
overturning failure occur, as the applied causative of 12MN (mass of 1200Mg). It is founded with a surface
moment quickly reverses, and a similar bearing-capacity square foundation (side B) on a 25m thick stiff clay deposit
mechanism may develop under the other edge. The (Su = 150kPa).
problem again reduces to computing the inelastic
deformations, i.e. permanent rotation.

Figure 6.8: Computed displacement response time-histories for the two designs of Figure 6, for the “small” intensity
(Kalamata 1986) motion.

Figure 6.9: Computed moment−rotation and settlement−rotation hysteretic response of the two foundations of Figure 6.6,
for the “small” intensity (Kalamata 1986) motion.
2924 B. Simpson et al. / State of the art Report: Analysis and Design

Two foundation solutions are examined: b. B = 7 m. This corresponds to the new design concept,
a. B = 11m. This derives from a conventional slightly- where significant plastic deformation is allowed to take
conservative design, which leads to a large static place in the foundation–soil system, to the point of
bearing−capacity factor of safety (FSV ≈ 5.6) and a mobilisation of the bearing–capacity failure mechanisms.
pseudo-static bearing-capacity factor FSE ≈ 2, for a code- These may develop alternatingly on either side under the
specified design spectrum having A = 0.24g (horizontal), footing, as large cyclic overturning moments arise during
corresponding to soil category “B”, and an estimated shaking. This design is barely adequate under static
“behaviour” or “q” factor of 2. Although the value of 2 vertical loads (FSV ≈ 2.8); under the design earthquake it
is high for a seismic FS, it was chosen in the anticipation leads to a pseudo-static FSE = 0.50 ⎯ well below unity to
of much-stronger acceleration histories than those be acceptable within conventional engineering thinking.
specified by the A = 0.24g code spectrum. It is thus quite It is therefore expected that during shaking by a
possible that during a design or stronger–than–design design−level, and especially by an above−design−level,
event structural (bending) plastic hinging will develop at ground motion the soil “failure” mechanisms will
the base of the column, with a rather minimal inelastic develop. The question is what the consequences will be
action in the soil or the soil–footing interface ⎯ a for the foundation and the superstructure, and how the
typically prudent conventional design. computed responses of the two systems of Fig. 6.6 differ
from one-another.

Figure 6.10: Computed displacement response time-histories for the two designs of Figure 6, for the “high”
intensity (Takatori 1995) motion.

Figure 6.11: Computed moment−rotation and settlement−rotation hysteretic response of the two
foundations of Figure 6.6, for the “high” intensity (Takatori 1995) motion.
B. Simpson et al. / State of the art Report: Analysis and Design 2925

This is demonstrated with Figures 6.8 - 6.11. Two be very small ⎯ for this particular excitation, undoubtedly
different real accelerograms are used as excitation: coincidentally, they almost vanished!
• The Kalamata Administration Building record of the 1986 In conclusion, inelastic “failure” mechanisms could be
Ms = 6.2 earthquake (Gazetas et al 1991). With an A ≈ allowed to develop in soil-foundation systems designed for
PGA ≈ 0.26g but a response spectrum with values smaller strong seismic excitation. Mobilisation of such mechanisms
than those of the design code spectrum at the period range does not usually lead to failure; it may in fact prove quite a
of interest, this motion will be referred to as “Small” beneficial way to save the structure.
Intensity excitation. Two supporting case histories are mentioned here: (a) the
settlement of slender buildings in Adapazari during the 1999
• The Takatori record of the 1995 MJMA = 7.2 Kobe
Kocaeli Earthquake, despite mobilisation of bearing-capacity
Earthquake, which in addition to its high PGA, 0.63 g,
“failure” mechanisms (as tentatively explained by Gazetas et
has spectral values in excess of 1.5g over a very wide al 2003). (b) the behaviour of the Kobe harbour breakwaters
period range (0.30 sec – 1.20 sec) . It thus undoubtedly with no overturning and only small permanent lateral
constitutes a “High” Intensity excitation ⎯ substantially displacements and rotations; whereas by contrast, the uni-
larger than a design excitation. directionally moving soil-supporting quay walls (caissons
Figures 6.8 and 6.9 show the results for the “Small” identical to the breakwaters) suffered huge seaward
Intensity shaking. Specifically, Figure 6.8 plots the time displacements and rotations (Sekiguchi et al 1996, Dakoulas
histories of the main components of superstructure and Gazetas 2007).
displacement:
• the displacement of superstructure mass due to foundation
rocking: uș (in blue) 7 CONCLUDING REMARKS
• the (additional) displacement of the superstructure,
representing distortion of the column due to bending: ustr This paper has presented highlights from recent developments
(grey) in some important areas of geotechnical analysis and design.
• the algebraic sum of the above two: utot = ustr + uș Whilst analysis plays an important role, the process of design
(black) is wider, incorporating empirical methods based on carefully
The lateral displacement of the foundation is not shown, as recorded experience, together with adequate factors or
it is quite secondary for the particular chosen slender system. margins to provide society with the level of safety and
Figure 6.9 refers to the response of the foundation. It plots serviceability recognised as needed on the basis of experience
the dynamic overturning moment-rotation, M − ș, relation and tradition. Provision of adequate contractual processes are
and the dynamic settlement–rotation, w−ș, relation for each also an integral part of design.
of the two foundations. The conclusions emerging from the In the development of codes and standards, attempts are
figures are clear: being made to rationalise the safety processes, though
reference to past practice is still deemed to be irreplaceable.
• The conventional foundation (B = 11 m) experiences very
Similarly, the valuable results of advanced analytical
small nearly-elastic rotation, with a small accumulation of
techniques must be reviewed and interpreted in the light of a
cyclic settlement (≈ 2.5 cm). By contrast, the column thorough understanding of the real behaviour of structures
experiences large distortion, almost 10 cm, as if the interacting with the ground.
structure were responding on a fixed base. The authors hope that pooling experience from different
• The “daring” foundation design, B = 7 m, experiences continents, varying geologies, onshore and offshore will
very large rotation, with uș reaching 12 cm. The enable geotechnical engineers to provide more reliable
significant inelastic action in the soil is reflected in the analysis and to derive sound designs.
highly-hysteretic M − ș relationship, as well as in an
appreciable accumulated foundation settlement ( ≈ 6 cm).
By refreshing contrast, the structural distortion has been ACKNOWLEDGEMENTS
limited to merely 2 cm ⎯ indicative of almost elastic
column response. Dr Simpson and Dr Morrison gratefully acknowledge the
In similar fashion, Figures 6.10–6.11 compare the results technical and financial support of Arup Geotechnics, in
for the “High” Intensity (Takatori) excitation. The following particular the help of Jon Shillibeer and Kelly Byrne.
is a summary of important conclusions from these plots: Valuable advice has been provided by Professor Roger Green,
• The conventional foundation design, B = 11 m, with too Dr Dennis Becker of Golder Associates, Professor KK Phoon
little help from the (“unyielding”) foundation, cannot of National University of Singapore and Dr Masahiro Shirato,
cope with the huge accelerations of the Takatori record. CAESAR, PWRI, Japan
mechanisms. The maximum rotation șmax reaches 0.036 The work reported by Professor Gazetas was conducted
rad, corresponding to a substantial deck displacement uș for the “DARE” project, financially supported by a European
≈ 50 cm. Accumulated settlement: wmax = wres ≈ 26 Research Council (ERC) Advanced Grant under the “Ideas”
cm. The superstructure, however, hardly deforms and Programme in Support of Frontier Research (Grant
thus remains safe despite the much-larger-than-design Agreement 228254).
ground shaking. With a warning: the above significant
displacements (26 cm and 50 cm) may imply such large
differential settlements between adjacent foundations and REFERENCES
differential displacements between adjacent piers that
AASHTO. 2008. LRFD Bridge Design Specifications, 4th edition,
indirect structural damage is unavoidable. Note also that 2007 with 2008 Interim Revisions. American Association of
the developing accelerations in the structure are also State Highway and Transportation Officials.
smaller when the soil is yielding and the foundation ACI 224R-01. 2008. American Concrete Institute 224R – Control of
uplifts ⎯ an additional benefit effect from foundation Cracking in Concrete Structures.
plastic “hinging”. ACI 318. 2008. American Concrete Institute 318 - Building Code
One might arguably consider the above footing−related Requirements for Structural Concrete.
deformations as excessive. Notice, however, that these were American Association of State Highway and Transportation Officials
peak values; the residual rotation and displacement appear to (AASHTO). 2002. Standard Specifications for Highway Bridges.
2926 B. Simpson et al. / State of the art Report: Analysis and Design

American Railway Engineering and Maintenance-of-Way Bray, J.D., Seed, R.B., & Seed, H.B. 1994b. Analysis of Earthquake
Association (AREMA). 2008. Manual for Railway Engineering. Fault Rupture Propagation through Cohesive Soil. Journal of
Anastasopoulos, I. Bransby, F. 2008b. Design of Slab Foundations Geotechnical Engineering, ASCE, Vol. 120, No.3, March, pp.
against Seismic Faulting. Part 2: Soil – Structure Interaction. 562-580.
Proc. 3rd National Conference of Earthquake Engineering and Brinch Hansen, J. 1961. The ultimate resistance of rigid piles against
Engineering Seismology, Athens, 5–7 November, Paper 1958. transversal forces. Danish Geotechnical Institute Bulletin No.
Anastasopoulos, I. Gazetas, G. Drosos, V. Georgarakos, T. & 12.
Kourkoulis, R.. 2008a. Design of bridges against large tectonic BS5400. 1988. Steel, concrete and composite bridges. BSI, London.
deformation. Earthquake Engineeing and Engineering Vibration, BS8002. 1994. British Standards Institution Code of Practice for
Vol. 7, pp. 345–368. Earth Retaining Structures.
Anastasopoulos, I. Gerolymos, N. Gazetas, G. & Bransby, M. F. BS8110. 1997. Structural use of concrete. BSI, London.
2008b. Simplified approach for design of raft foundations BSI. 2004. BS EN 1997:1: 2004. Eurocode 7: Geotechnical design -
against fault rupture. Part I: Free-field. Earthquake Engineering Part 1: General rules.
and Engineering Vibration, Vol. 7, pp. 147–163. BSI. 2007. UK National Annex to Eurocode 7: Geotechnical design -
Anastasopoulos, I. Callerio, A. Bransby, M.F. Davies, M.C.R., El. Part 1: General rules.
Nahas, A. Faccioli, E. Gazetas, G. Masella, A. Paolucci, R. BSI. 2007a. BS EN 1997–2:2007 Eurocode 7 Geotechnical design,
Pecker, A. & Rossignol, E. 2008c. Numerical Analyses of Fault– Part 2: Ground Investigation and testing.
Foundation Interaction. Bulletin of Earthquake Engineering, BSI. 2007b. UK National Annex to Eurocode 7: Geotechnical design
Special Issue: Integrated approach to fault rupture- and soil- - Part 1: General rules.
foundation interaction , Vol. 6, pp. 645–675. Burland, J.B, Potts, D.M. &Walsh, N.M. 1981. The overall stability
Anastasopoulos, I. Gazetas, G. Bransby, F. Davies, M.C.R. & Nahas, of free and propped embedded cantilever walls. Ground
E. A. 2009. Normal Fault Rupture Interaction with Strip Engineering 14(5), 28-38.
Foundations. Journal of Geotechnical and Geoenvironmental Butcher AP, Powell, J.J.M & Skinner, H.D. 2006a. Reuse of
Engineering, American Society of Civil Engineers , Vol. 135, No Foundations for Urban sites: A best practice handbook.
3, pp. 359-370. IHS/BRE press.
Anastasopoulos, ǿ. 2005. Fault Rupture–Soil–Foundation–Structure Butcher AP, Powell, J.J.M & Skinner, H.D (editors), 2006b.
Interaction, Ph.D. Dissertation, School of Civil Engineering, Proceedings of International Conference on Reuse of
National Technical University. Athens, pp.570. Foundations for Urban sites. IHS/BRE press.
API RP 2A-WSD. 2000. Recommended practice for planning, Cathart, A. 2003. Channel Tunnel Rail Link: A Contract Partnership.
designing, and constructing fixed offshore platforms - Working Civil Engineering Vol. 156, Issue 5, pp 41-44.
stress design. - 21st edition. Central Electricity Generating Board. 1965. Report of the Committee
Apostolou, M. Gazetas, G. & Garini, E. 2007. Seismic response of of Inquiry into Collapse of Cooling Towers at Ferrybridge
slender rigid structures with foundation uplifting. Soil Dynamics Monday 1 November 1965. Central Electricity Generating
and Earthquake Engineering (in press). Board, London.
Asian Technical Committee No.3 (ATC3). 1997. Manual for Chapman, T. Connolly, M. Nicholson, D. Raison, C & Yeow. H.
zonation on areas susceptible to rain-induced slope failure, 81p. 1999. Advances in understanding of base grouted pile
Barnes, I. & Sear, L. 2004. Ash Utilisation from Coal-Based Power performance in very dense sand. International symposium and
Plants. Report No. COAL R274, DTI/Pub URN 04/1915. UK exhibition in tunneling construction and piling, London, 1999, pp
Department of Trade and Industry. 57-69.
Becker, D.E. 1996. Eighteenth Canadian Geotechnical Colloquium: Chau, C. Nicholson, D. & Soga, K. 2006. Comparison of embodied
Limit states design of foundations. I: An overview of the energy of four different retaining wall systems. Proc. Int. Conf.
foundation design process. Canadian Geotech J, 33(6), 956-983. Reuse Found. for Urban Sites (RuFus 2006), 277–285.
Becker, D.E. 2008. Personal communication. Cheng, Y.M. 2004. Géotechnique 54, No 2, 149-150.
Berezantzev, V.C. Khristoforov, V. & Golubkov, V. 1961. Load Clausen, C.J.F. Aas, P.M & Karlsrund., K. 2005. Bearing capacity of
bearing capacity and deformation of piled foundations. Proc 5th driven piles in sand, the NGI approach. Int. Symp. on Frontiers
Int. Conf. Soil Mech. Found. Engng., Paris., 11-15. in Offshore Geotechnics, Perth, Australia (ISFOG 2005).
Berill, J.B. 1983. Two-dimensional analysis of the effect of fault Cremer, C., Pecker, A. & Davenne, L. 2002. Modelling of nonlinear
rupture on buildings with shallow foundations. Soil Dynamics dynamic behaviour of a shallow strip foundation with macro-
and Earthquake Engineering, Vol. 2, No. 3, pp. 156-160. element. Journal of Earthquake Engineering 6, No. 2, 175-211.
Bilfinger, W. 2005. Design Procedure for Groundwater Loads on Dahlberg,R and Ronold, KO (1993) Limit state design of offshore
Tunnel Linings., Symposium on Waterproofing of Underground foundations. Proc Int Symp Limit state design in geotechnical
Structures, San Paulo Brazil. engineering, Vol 2, pp491-500. Danish Geotechnical Society.
Bolton, M.D. 1987. The strength and dilatancy of sands. Dakoulas, P. & Gazetas, G. 2008. Insight to Seismic Earth and Water
Géotechnique, 37, No. 2, 219-226. Pressures Against Caisson Quay Walls. Geotechnique, Vol. 58,
Bolton, M.D. 1993. What are partial factors for? In Proceedings of No. 2, pp 95-113.
the International Symposium on Limit State Design in Dansk Ingeniorforening. 1984. Danish Code of Practice for
Geotechnical Engineering, Copenhagen, May 26-28, 1993. Foundations Engineering (Dansk Standard DS 415). English
Sponsored by the Danish Geotechnical Society. Vol. 3, pp. 565- translation.
583. Det Norske Veritas. 1993. Rules for classification of fixed offshore
Bond, A. & Harris, A. 2008. Decoding Eurocode 7. Taylor and installations.
Francis, London. Duddeck, H. & Erdmann, J. 1982. Structural design models for
Bond, A.J. & Simpson, B. 2009. Pile design to Eurocode 7 and the tunnels. Tunnelling 82, pp 83-91.
UK National Annex. Submitted to Geotechnical Engineering. EN1990. 2002. Eurocode: Basis of design. BSI, London. (BS EN
Thomas Telford. 1990:2002).
Boverket. 1995. Design Regulations BKR 94 – Mandatory provisions EN1992. 2004. Eurocode 2 – Design of concrete structures. BSI,
and general recommendations. Swedish Board of Housing, London. (BS EN 1992:2004).
Building and Planning. EN1997. 2004. Eurocode7: Geotechnical design. BSI, London. (BS
Brandl, H. 2006. Energy Foundations and other thermo-active EN 1997:2004).
ground structures. Géotechnique 56, No 2, 81-110. EN1997-1. 2004. Eurocode7: Geotechnical design - Part 1: General
Bray, J.D. 1990.). The effects of tectonic movements on stresses and rules. BSI, London. (“EC7”).
deformations in earth embankments. Ph.D. Dissertation, Eurocode EC8. 2000. Structures in Seismic Regions. Part 5:
University of California, Berkeley. Foundations, retaining structures, Geotechnical Aspects.
Bray, J.D., Ashmawy, A., Mukhopadhyay, G., Gath E.M. 1993. Use Faccioli, E. Anastasopoulos, I. Callerio, A. & Gazetas, G. 2008. Case
of Geosynthetics to Mitigate Earthquake Fault Rupture histories of fault–foundation interaction. Bulletin of Earthquake
Propagation Through Compacted Fill. Proceedings of the Engineering, Special Issue: Integrated approach to fault rupture-
Geosynthetics ’93 Conference, Vol. 1, pp. 379-392. and soil-foundation interaction, Vol. 6, pp. 557–583.
B. Simpson et al. / State of the art Report: Analysis and Design 2927
FEMA 356. 2000: Prestandard and Commentary for the Seismic Johansson, J. & Konagai, K. 2004. Fault induced permanent ground
Rehabilitation of Buildings. deformations–Simulations and experimental verification.
Fleming, W.G.K, Weltman, A.J, Randolph, M.F & Elson, W.K. Proceedings of the 13th World Conference on Earthquake
1992. Piling engineering. end edn. WIley. Engineering, August 1–6, Vancouver, Canada.
Fleming, W.G.K, Weltman, A.J, Randolph, M.F & Elson, W.K. Johnston, I.W. & Lam T.S.K. 1989. Shear behaviour of regular
2008. Piling engineering. 3rd edition. Taylor and Francis Ltd. triangular concrete/rock joints – analysis. Journal of
Foye, K.C, Salgada, R. & Scott, B. 2006a. Assessment of variable Geotechnical Engineering, ASCE, Vol 115.5, pp 711-725.
uncertainties for reliability-based design of foundations. J Kanto Branch of the Japanese Geotechnical Society. 2007. Guideline
Geotech Geoenviron Eng, 132(9), 1197-1207. for site investigation, analyses and countermeasures of existing
Foye, K.C. Salgado, R. & Scott, B. 2006b. Resistance factors for use reclaimed housing lots against earthquakes, 201p. (in Japanese)
in shallow foundation LRFD. J. Geotech. Geoenviron. Eng., Kawashima, K. & Hosoiri, K. 2005. Rocking Isolation of Bridge
132(9), 1208–1218. Columns on Direct Foundations. Journal of Earthquake
Fugro. 2004. Fugro Engineers B.V., Axial pile capacity design Engineering, 6(2).
method for offshore driven piles in sand. Klotz, E.U. & Coop, M.R. 2001. An investigation of the effect of soil
Gaba, A.R, Simpson, B. Powrie, W. & Beadman, D.R. 2003. state on capacity of driven piles in sands. Géotechnique 51, No.
Embedded retaining walls: guidance for economic design. 9, 733-751.
CIRIA Report C580. Lambe, T.W & Whitman, R.V. 1979. Soil mechanics, SI version.
Gajan, S. Kutter, B. Phalen, J. Hutchinson, T. & Martin, G. 2005. Wiley.
Centrifuge modelling of load-deformation behaviour of rocking Lazarte, C.A. & Bray, J.D. 1995. Observed Surface Breakage due to
shallow foundations. Soil Dynamics and Earthquake Strike-Slip Faulting, Third International Conference on Recent
Engineering, (in press). Advances in Geotechnical Engineering and Soil Dynamics, Vol.
Gazetas, G. Anastasopoulos, I. Apostolou, M. 2007. Shallow and 2, pp. 635–640.
Deep Foundations Under Fault Rupture or Strong Seismic Lehane, B.M. & Randolph, M.F. 2002. Evaluation of a minimum
Shaking. Geotechnical Earthquake Engineering, Springer, base resistance for driven piles in siliceous sand. Journal of
Chapter 9, pp. 185-217. Geotechnical and Geoenvironmental Engineering, ASCE, Vol.
Gazetas, G. Dakoulas, P. & Papageorgiou, A. 1990. Local-Soil and 130.6, pp 656-658.
Source Mechanism Effects in the 1986 Kalamata Lehane, B.M. Schneider, J.A. & Xu, X. 2005a. CPT based Design of
Earthquake. Earthquake Engineering & Structural Dynamics, Driven Piles in Sand for offshore structures. University of
Vol. 19, pp. 431-456. Western Australia report Geo 05345.
Gazetas, G. Pecker, A. Faccioli, E. Paolucci, R. & Anastasopoulos, I. Lehane, B.M. Schneider, J.A. & Xu, X. 2005b. The UXA-05 method
2008. Design Recommendations for Fault–Foundation for prediction of axial capacity of driven piles in sand. Frontiers
Interaction. Bulletin of Earthquake Engineering, Special Issue: in Offshore Geotechnics: ISFOG 2005, Taylor and Francis
Integrated approach to fault rupture- and soil-foundation Group, pp 683-689.
interaction, Vol. 6, 677–687. Lord, J.A, Clayton, C.R.I. & Mortimore, R.N. 2002. Engineering in
Gazetas, G. & Apostolou, M. 2004. Nonlinear soil-structure Chalk. Ciria Report No C574.
interaction: foundation uplift and soil yielding. Proc. 3rd US- Lyamin, A. V. Salgado, R. Sloan, S. W. & Prezzi, M. 2006. Three
Japan Workshop on Soil-Structure Interaction, USGS, M. dimensional bearing capacity of footings in sand. Working
Todorovska and M. Celebi editors, Menlo Park, CA. paper, Purdue Univ.
GEO. 2006. Foundations design and Construction. Geo Publication //http://www.ecn.purdue.edu/~rodrigo/papers/.
1/2006, The Government of Hong Kong. Magnus, R. Teh, C.I. & Lau, J.M. 2005. Report on the Incident at the
Green, R. 2008. Personal communication. MRT Circle Line worksite that led to the collapse of the Nicoll
Honjo, Y.Y. Kikuchi, M. Suzuki, K. Tani & M. Shirato. 2005. JGS Highway on 20 April 2004. Subordinate Courts, Singapore.
Comprehensive Foundation Design Code: Geo-code 21, Proc. Makris, N. & Roussos, Y.S. 2000. Rocking Response of Rigid Blocks
16th ICSMGE, pp.2813-2816, Osaka. under Near-Source Ground Motions. Geotechnique, 50(3), 243-
Housner, G. 1963. The Behavior of Inverted Pendulum Structures 262.
During Earthquakes. Bulletin of the Seismological Society of Mandolini, A. Russo, G. & Viggiani, C. 2005. Pile foundations:
America. 53(2), 404-417. experimental investigation, analysis and design. State of the Art
Hovarth, R.G. Kenny, T.C. & Trow, W.A. 1980. Results of tests to Report. Proc. 16th International Conference on Soil Mechanics
determine shaft resistance of rock socketed drilled piers. Int. and Geotechnical Engineering, Osaka. Vol. 1, 177-213.
conf. on structural foundations on rock, Sydney. Martin, C.M. & Randolph, M. F. 2006. Upper-bound analysis of
Iai, S. Matsunaga, Y. & Kameoka, T. 1992. “Strain space plasticity lateral pile capacity in cohesive soil. Géotechnique, Vol. 56.2,
model for cyclic mobility,” Soils and Foundations, Vol.32, No.2, pp 141-145.
pp.1-15, Ishihara, K. and Yasuda, S. 1975. Undrained deformation Matsuo, M. and Kawamura, K. 1977. Diagram for construction
and liquefaction of sand under cyclic stresses, Soils and Foundations, control of embankment on soft ground. Soils and Foundations,
Vol.23, No.6, pp.29-35. Vol.17, No3, pp37-52.
Iwasaki, T. Tatsuoka, F. Tokida, K. & Yasuda, S. 1978. A prediction Mayne, P.W. & Kulhawy, F.H. 1982. Ko – OCR relationship in soil.
method for assessing soil liquefaction potential based on case Geotechnical Engineering, ASCE. Vol 108.6; 851-872
studies at various sites in Japan, Proc. of the 2nd Int. Conf. on Mehta, P.K. 2001. Reducing the Environmental Impact of Concrete.
Microzonation, Vol.2, pp.885-896. Concrete International, October 2001.
Japan Road Association. 1986. Earthwork manual for soft ground. Mehta, P.K. 2004. High Performance, High Volume Fly Ash
(in Japanese). Concrete For Sustainable Development Proc Int Workshop on
Japanese Geotechnical Society. 2006. Japanese Geocode 21: Sustainable Development and Concrete Technology, Beijing
Principles for Foundation Designs Grounded on a Performance- Iowa State University.
based Design Concept, JGS 4001-2004. http://www.ctre.iastate.edu/pubs/sustainable/index.htm.
Japanese Geotechnical Society. 1998. Remedial measures against soil Meyer, C 2005. Concrete as a Green Building Material. Invited
liquefaction, 443p.Balkema. Lecture, Proceedings of the Third Int. Conference on
Japanese Geotechnical Society. 2004. Reconnaissance report on the Construction Materials, ConMat'05, Vancouver, Aug. 22-25.
2003 Tokachi-oki earthquake. (in Japanese). Monzon, J.C. 2006. Review of CPT based design methods for
Japanese River Association. 1997. “Manual criteria for river works,” estimating axial capacity of driven piles in siliceous sand.
(in Japanese). M.Eng thesis submitted to MIT, USA.
Japanese Society of Soil Mechanics and Foundation Engineering. Nakai, T. Muir Wood, D. & Stone, K.J.L. 1995. Numerical
1994. Reconnaissance report on the 1993 Kushiro-oki Calculations of Soil Response Over a Displacing Basement. Soils
earthquake and the 1993 Notohaoto-oki earthquake. (in and Foundations, Vol. 35, No.2, pp. 25–35.
Japanese). Nakamura, Y. & Murakami, Y. 1980. A study on the seismic stability
Jardine, R. Chow, F. Overy, R. & Standing, J. 2005. ICP Design of embankments of three big rivers in Nobi Plain, 34th Technical
Methods for Driven Piles in Sands and Clays. Thomas Telford, Conference of Ministry of Construction, pp.96-104. (in
London. Japanese).
2928 B. Simpson et al. / State of the art Report: Analysis and Design

Newmark, N. M. 1959. A method of computation for structural Roth, W.H. Scott, R.F. & Austin, I. 1981. Centrifuge modelling of
dynamics, ASCE, Vol.85, EM3, pp.67-94. fault propagation through alluvial soils. Geophysical Research
Nicholson, D Chapman, T & Morrison, P. 2002. Pressuremeter Letters, Vol. 8, No. 6, pp. 561–564.
proves its worth in London’s Docklands. Ground Engineering Rowe, R.K. & Armitage, H.H. 1987. A design method for drilled
March 2002, pp 32-34. piers in soft rock. Canadian Geotechnical Journal, Vol 13.3, pp
Nicholson, D.P. Dew, C.E. & Grose, W.J. 2006. A Systematic ‘Best 324-333.
Way Out’ Approach using Back Analysis and the Principles of Salençon, J. & Pecker, A. 1995. Ultimate bearing capacity of shallow
the Observational Method. International Conf on Deep foundations under inclined and eccentric loads. Part II: Purely
Excavations, 28-30 June 2006, Singapore. cohesive soil without tensile strength. Eur. J. Mech. A/Solids 14,
Oka, F. Yashima, A. Tateishi, A. Taguchi, Y. & Yamashita, S. 1999. No. 3, 377-396.
“A cyclic elasto-plastic constitutive model for sand considering a Sasaki, Y. Kano, S. & Matsuo, O. 2004. Research and practices on
plastic-strain dependence of the shear modulus.” Geotechnique remedial measures for river dikes against soil liquefaction,
Vol.49, No. 5, pp.661-680. Journal of Japan Association for Earthquake Engineering, Vol.4,
Orr, T.L.L. 2005. Model Solutions for Eurocode 7 Workshop No.3 (Special Issue), pp.312-335.
Examples. Trinity College, Dublin. Sawada, R. Tanamura, S. Nishimura, A. & Koseki, J. 1999.
Osman A.S. and Bolton M.D. 2006. Design of braced excavations to Simplified method to subsidence of embankment on liquefiable
limit ground movements. Proceedings of Institution of Civil ground. Proc. 34th Japan National Conference on Geotechnical
Engineers, Geotechnical Engineering 159 (3), 167-175. Engineering. pp.2091-2092. (in Japanese).
Osman, A.S. Yeow, H.C. Bolton, M.D. 2004. Estimation of Schneider, H. R. 1997. Definition and determination of characteristic
Undrained Settlement of Shallow Foundations on London Clay. - soil properties. Contribution to Discussion Session 2.3, XIV
First International Conference on Structural and Foundation ICSMFE, Hamburg. Balkema.
Failures August 2-47-9, 2004, Taipei, Taiwan Singapore. Salgado, R. Lyamin, A. V. Sloan, S. W. & Yu, H. S. 2004. Two- and
Padfield, C.J. & Mair, R.J. 1984. Design of retaining walls embedded three-dimensional bearing capacity of foundations in clay.
in stiff clay. CIRIA Report 104. Géotechnique 54, No.5, 297-306.
Pecker A. 2005. Design and Construction of the foundations of the Schulze, H and Duddeck, H 1964. Spannungen in
Rion-Antirrion Bridge. Proceedings, 1st Greece Japan Workshop schildforgetriebenen Tunneln. Beton Stahlbetonb., 59.
on Seismic Design, Observation and Retrofit of Foundations Schuppener, B. Simpson, B. Orr, T.L.L. Frank, R. Bond, A.J. 2009.
Pells, P.J.N., Rowe, R.K., & Turner, R.M. 1980. An experimental Loss of static equilibrium of a structure - definition and
investigation into side shear for socketed piles in sandstone. In verification of limit state EQU. 2nd Int Symp on Geotechnical
Proceedings of the International Conference on Structural Safety and Risk, Gifu.
Foundations on Rock, Sydney, pp. 291–302. Seidel, J.P. & Collingwood, B. 2001. A new rock socket roughness
Pietruszezak, S.T., & Mroz, Z. 1981. Finite element analysis of factor for prediction of rock socket shaft resistance. Canadian
deformation of strain softening materials. International Journal Geotechnical Journal, Vol. 38, pp 138-153.
for Numerical Methods in Engineering, Vol. 17, pp. 327–334. Seifart, R. 2007. Permafrost, A building problem in Alaska.
Pollalis, S.N. 2006. Channel Tunnel Rail Link, Risk Transfer and Published by University of Alaska Fairbanks.
Innovation in Project Delivery., Harvard Design School, Shirato, M. 2008. Personal communication.
November 2006. Sidel, J. P. & Haberfield, C.M. 1995a. Towards an understanding of
Potts, D. M. Dounias, G. T. & Vaughan, P. R. V. 1987. Finite joint roughness. Int Journal of Rock Mechanics and Rock
element analysis of the direct shear box test. Géotechnique, Vol. Engineering, Vol. 28.2, pp 69-92.
37, No. 1, pp. 11–23. Sidel, J. P. & Haberfield, C.M. 1995b. The axial capacity of pile
Poulos, H. G. & Davids, J. 2005.Foundation design for the Emirates sockets in rocks and hard soils. Ground Engineering, Vol. 28.2,
Twin Towers, Dubai. Canadian Geotechnical Journal, Vol 42, pp pp 33-38.
716-730. Simpson, B. (2007) Approaches to ULS design - The merits of Design
Poulos, H.G. & Davis E. 1980. Pile Foundation Analysis and Approach 1 in Eurocode 7. ISGSR2007 First International
Design. Symposium on Geotechnical Safety & Risk pp 527-538.
Pristley, N. Seible, F. & Calvi, G.M. 1996. Seismic Design and Shanghai Tongji University, China.
Retrofit of Bridges. John Wiley and Sons, Inc., New York. Simpson, B. & Driscoll, R. 1998. Eurocode 7 - a commentary.
Public Works Research Institute. 1997. Design and construction Construction Research Communications Ltd, Watford, UK.
manual for remediation of river dikes against liquefaction, Simpson, B. Nicholson, D.P. Banfi, M. Grose, W.G. & Davies, R.V.
Research report of PWRI, Vol.3513. (in Japanese). 2008).Collapse of the Nicoll Highway excavation, Singapore.
Railway Technical Institute. 1999. Seismic design code for Japanese Proc Fourth International Forensic Engineering Conference.
railway facilities, (in Japanese). Thomas Telford.
Randolph, M.F. 2003a. Science and empiricism in pile foundation Simpson, B. Yeow, H.C. Pillai, A.K. & Grose, W.J. 2006. Benefits
design. Géotechnique Vol. 53.10 847-875. derived from use of 3D finite element analysis in the design of
Randolph, M. F. 2003b. PIGLET, Analysis and design of Pile deep excavations and tunnels. Proc Int Conf on Deep
Groups, Version 4.2. (User manual). Excavations 28-30 June 2006, Singapore.
Randolph, M. Cassidy, M. Gourvenec, S. & Erbich, C. 2005. Technical Committee on the 1968 Tokachi-oki earthquake. 1969.
Challenges of Offshore Geotechnical Engineering. State of the Reconnaissance report on the 1968 Tokachi-oki earthquake. (in
Art Report. Proc. 16th International Conference on Soil Japanese).
Mechanics and Geotechnical Engineering, Osaka. Vol. 1, 123- Tohoku Branch, JSCE. 1980. Reconnaissance report on the 1978
176. Miyagiken-oki earthquake. (in Japanese).
Reece, L.C. 1997. Analysis of laterally loaded piles in weak rock. Transportation Research Board. 2007. Guidelines for Using
Journal of Geotechnical and Geoenvironmental Engineering, Environment–Friendly Concrete (EFC) in Transportation
Vol. 123.11, pp 1010-1017. Infrastructure, Sponsoring Committee: AFN10, Basic Research
River Bureau. 1995. Technical manual for seismic diagnosis of and Emerging Technologies Related to Concrete.
existing river dikes, River Bureau, MLIT. (in Japanese) Tsuchida, T. Sato, T. Hong, Z. Mitsukuri, K. & Sakai, K. 2001. Field
River Bureau. 2007. Performance-based seismic design criteria for Placing Test of Lightweight Treated Soil at 10m Seawater Depth
river facilities (Draft). (in Japanese). in Kumamoto Port. Technical Note of the Port and Airport
Rock Engineering. 1994. Engineering and Design – Rock Research Institute, no. 1007. http://sciencelinks.jp/j-
Engineering. US Army Corps of Engineers, Report No EM east/article/200204/000020020402A0106434.php
1110-1-2908 (http://140.194.76.129/publications/eng- Turner, J. 2006. Rock-Socketed shafts for Highway structure
manuals/em1110-1-2908/toc.htm). foundations – A Synthersis for Highway Practice. Transport
Rostam, S. & Høj, N.P. 2004. Integrated Tunnel Design from Cradle Research Board, Washington, USA.
to Grave, Enhancing Structural Performance and Owner US Army. 1994. Engineering and Design - Rock Foundations, US
Confidence. Safe & Reliable Tunnels. Innovative European Army Corps of Engineers, EM 1110-1-2908.
Achievements, First International Symposium, Prague 2004.
B. Simpson et al. / State of the art Report: Analysis and Design 2929
Ventouras, K. & Coop, M. R. 2009. On the behaviour of Thanet Yasuda, S. Hitomi, T. & Hashimoto, T. 2004. A detailed study on the
Sand: an example of an uncemented natural sand. liquefaction-induced settlement of timber houses during the 2000
Géotechnique, in press. Tottoriken-seibu earthquake, Proc. of the 5th International
Vogt, N. & Schuppener, B. 2006. Implementation of Eurocode 7-1 Conference on Case Histories in Geotechnical Engineering,
Geotechnical design in Germany. Proc TAIPEI2006 International Paper No.3.33.
Symposium on New Generation Design Codes for Geotechnical Yasuda, S. Ideno,T. Sakurai, Y. Yoshida, N. & Kiku, H. 2003.
Engineering Practice Nov. 2~3, 2006, Taipei, Taiwan. Analyses for liquefaction-induced settlement of river levees by
Walters, J.V. & Thomas, J.N. 1982. Shear Zone Development In ALID. Proc. 12th Asian Regional Conference on Soil Mechanics
Granular Materials. Proceedings of the 4th International and Geotechnical Engineering, pp.347-350.
Conference on Numerical Methods in Geomechanics, Edomnton, Yasuda, S. Ideno, T. Sakurai, Y. Yoshida, N. & Kiku, H. 2003.
Canada, May 31 – June 4, Vol. I, pp. 263-274. Analyses for liquefaction-induced settlement of river levees by
White, D.J. & Bolton, M.D. 2005. Comparing CPT and pile base ALID, Proc. of the 12th Asian Regional Conference on Soil
resistance in sand. Proc. Inst. Civil Eng. Geotechnical Mechanics and Geotechnical Engineering, pp.347-350.
Engineering158, Jan, 2005, 3-14. Yasuda, S. Morimoto, I. Kiku, H. & Tanaka, T. 2004.
Xu, X. & Lehane, B.M. 2005. Evaluation of end-bearing capacity of Reconnaissance report on the damage caused by three Japanese
closed ended pile in sand from cone penetration data. Proc earthquakes in 2003, Proc. of the 3rd International Conference
ISFOG 2005, Perth, Australia. on Earthquake Geotechnical Engineering, Vol.1, pp.14-21.
Yasuda, S. 2005. Recent several studies and codes on performance- Yazdchi, M. Macklin, S.R. Yeow, H.C. 2005. 3D modelling of
based design for liquefaction in Japan, Proc. of Geotechnical sprayed-concrete-lined tunnels in clay. Journal of Geotechnical
Earthquake Engineering Satellite Conference, 16th ICSMGE., Engineering, 159, 4, 243-250. Thomas Telford.
pp. 46-53. Yeow, H.C. Nicholson, D. Morrison, P. Coupland, J. & Kassir, M.
Yasuda, S. 2007. Countermeasures against seismic failure of existing 2005. Moorhouse, London – Designing for Future Crossrail.
reclaimed areas by banking. Keynote Lecture, Proc. of International DFI publication Deep Foundations, Winter 2006.
Workshop on Earthquake Hazards and Mitigations (EHAM 2007). Yeow, H.C. & Prust, R. 2005, "Feasibility of Modelling of Soil-
Yasuda, S. 2007. Methods for remediation of existing structures Structure Interaction Problems using Full Three-dimensional
against liquefaction, Proc. of the 4th International Conference on Finite Element Method", ASCE International Conference on
Earthquake Geotechnical Engineering, Keynote Lecture. Computing in Civil Engineering, Mexico.
Yasuda, S. Fujioka, K. Shiratori, S. & Mouri, T. 2008. Behaviour and Yeow, H.C. Wong, C. Pillai, A.K. & Simpson, B. 2005. The Use of a
Design of Expressway Embankments Subjected to Earthquakes, Substructure Method and Three-dimensional Finite Element
Proc. of the 3rd Taiwan-Japan Workshop on Geotechnical Modelling in Assessment of Damage due to Underground
Hazards for Large Earthquakes and Heavy Rainfalls, Keynote Tunnelling. 11th International Association of Computer Methods
Lecture, pp.1-20. and Advances in Geomechanics Conference, Italy.

Вам также может понравиться